id
stringlengths
27
33
source
stringclasses
1 value
format
stringclasses
1 value
text
stringlengths
13
1.81M
warning/0001/astro-ph0001276.html
ar5iv
text
# Implementing Feedback in Simulations of Galaxy Formation: A Survey of Methods ## 1. Introduction A detailed understanding of galaxy formation in hierarchically clustering universes remains one of the primary goals of modern cosmology. Whereas on large scales the clustering of matter is determined almost solely by gravitational forces, a large number of physical processes contribute to galaxy formation. Further complication is evident in that a significant number of these processes cannot be modelled from first principles in a simulation of galaxy formation, the main example of this being star formation. Analytic and semi-analytic theories of galaxy formation are well-developed (see White 1994, for an overview). The theoretical framework for studying the condensation of baryons in dark matter halos was laid out by White & Rees (1978) following the foundational work on hierarchical clustering by Peebles (1980, and references therein). White and Rees illuminated the fact that for baryons condensing in dark matter halos of sub-galactic ($`10^6`$ $`\mathrm{M}_{}`$) size the cooling time of the gas is always shorter than the free-fall collapse time. In a related paper, Fall & Efstathiou (1980) demonstrated that disk galaxies could be formed in a hierarchical clustering model, provided that the gas maintains its angular momentum. Later work by White & Frenk (1991) further developed this hierarchical clustering model and identified the cooling catastrophe for CDM cosmologies. The cooling catastrophe is caused by the large mass fraction in small high-density halos at early times. The gas resident in these halos cools on the time-scale of a Myr, thus precipitating massive star formation very early on in the development of the CDM cosmology (at odds with the observations of our Universe). To circumvent this problem White & Frenk introduced star formation and the associated feedback from supernovae and showed that, given plausible assumptions, the cooling catastrophe can be avoided. The main deficiency in the semi-analytic programmes is that they cannot describe the geometry of mergers which is exceptionally important in the assembly of galaxies. Smoothed Particle Hydrodynamic simulations of galaxy formation (Katz (1992); Navarro & White (1994); Evrard, Summers & Davis (1994), for example) detail the hierarchical merging history, but have been limited in terms of resolution. Achieving high resolution is particularly difficult. Since galaxy formation is affected by long range tidal forces, the simulation must be large enough to include these, which in turn enforces a low mass resolution. Two solutions to this problem exist; the first samples the long range fields using lower particle resolution than the main simulation region (the multiple mass technique, Porter 1985) while the second includes the long range fields as a pre-calculated low resolution external field (e.g. Sommer-Larsen, Gelato & Vedel (1998)). Simulations performed in this manner represent lengths scales from 50 Mpc to 1 kpc, a dynamic range of $`5\times 10^4`$. It is well known that in SPH simulations it is comparatively easy to form flattened disk structures resembling disk galaxies. However, the resulting gaseous structures are deficient in angular momentum (see Navarro, Frenk & White 1995, for a comparison of a number of simulations to observed galaxies). The loss of angular momentum occurs during the merger process as dense gas cores lose angular momentum to the dark matter halos (see Barnes 1992, for an explanation of the mechanism). This is a very significant problem since a fundamental requirement of the disk formation model, presented in Fall & Efstathiou (1980), is that the gas must maintain a similar specific angular momentum to the dark matter to form a disk. The solution to this problem is widely believed to be the inclusion of star formation and feedback from supernovae and stellar winds. By including these effects the gas should be kept in a more diffuse state which consequently does not suffer from the core-halo angular momentum transport problem. Notably, a recent letter by Dominguez-Tenreiro, Tissera & Saiz (1998), based upon analytic work by Christodoulou, Shlosman & Tohline (1995) and van den Bosch (1998), has shown that the inclusion of star formation may go some way in helping to resolve the angular momentum problem. They suggest that a second mechanism, bar formation, also contributes to the loss of angular momentum within the disk. Including a photoionizing background (Efstathiou 1992) appears to make the problem slightly less severe (Sommer-Larsen, Gelato & Vedel 1998) but this is the subject of debate; for a conflicting viewpoint see Navarro & Steinmetz (1997). Individual supernovae cannot be modelled in galaxy formation simulations. Consequently, there is no a priori theory for how feedback energy should be distributed in the simulation. A decision must be made whether feedback energy should be passed to the thermal or kinetic sector of the simulation. As a direct result of this ‘freedom’, a number of different algorithms for the distribution of feedback energy have been proposed (Katz (1992); Navarro & White (1993); Mihos & Hernquist (1994) for example). Since the interstellar medium is multi-phase (McKee & Ostriker (1977)), the feedback process should ideally be represented as an evaporation of molecular clouds. Note that in a single-phase model, the analogue of the cloud evaporation process is thermal heating. A seminal attempt at representing this process has been made by Yepes et al. (1997), and recently Hultman and Pharasyn (1999) have adapted the Yepes et al. model to SPH. At the moment it is unclear how well motivated these models are since (1) they must make a number of assumptions about the physics of the ISM and (2) they are not truly multiphase, since the dynamics are still treated using a single phase. Consequently, given the uncertainties inherent in multiphase modeling, the investigation presented here explores a single phase model. As a first approximation, the star formation algorithms can be divided into three groups. The first group contains algorithms which rely upon experimental laws to derive the star formation rate (Mihos & Hernquist (1994), for example). The second group contains those which predict the SFR from physical criteria (Katz (1992), for example). The third group is comprised of algorithms which do not attempt to predict the SFR, but instead set a density criterion for the gas so that when this limit is reached the gas is converted into stars (Gerritsen & Icke (1997), for example). In this work the first approach is taken. Given that feedback occurs on sub-resolution scales, it is difficult to decide upon the scale over which energy should be returned. However, SPH incorporates a minimum scale automaticallythe smoothing scale. Katz (1992) was the first to show that simply returning a specified amount of energy to the ISM is ineffective: the characteristic time-scale of radiative cooling at high density is far shorter than the simulation dynamical time-scale. This is, at least partially, another drawback of trying to model sub-resolution physics at the edge of the resolution scale. Further, the strong disparity between time-scales means that not only are length scales sub-resolution, so is the characteristic temporal evolution. This point has been made several times in relation to simulations with cooling (e.g. Katz 1992) but it seems even more important in the context of models with feedback. Thus, in an attempt to circumvent this problem, a new thermal feedback model is presented in which the radiative losses are reduced by changing the density used in the radiative cooling equation. The density used is predicted from the ideal gas equation of state and then integrated forward, decaying back to the SPH density (in a prescribed period). The effect of preventing radiative losses using a brief adiabatic period of evolution for the feedback region (Gerritsen (1997)) is also examined. An alternative method of returning thermal energy is to heat an individual SPH particle (Gerritsen (1997)). This method has been shown to be effective in simulations of isolated dwarf galaxies. The final mechanism considered is one that attempts to increase the energy input from SN to account for the high radiative losses. Mechanical feedback boosts are not considered for the following reasons; (1) parameters that are physically motivated (e.g. feedback efficiency of 10%) produce too much velocity dispersion, and (2) the method makes no account for force softening. The response of the (simulated) ISM to a single feedback event is examined to gain an understanding of the qualitative and quantitative performance of each feedback algorithm. To determine whether the parameters of the star formation are more important than the dynamics of the simulation, an exploration of the parameter space of one of the algorithms is undertaken. Since feedback is expected to have a more significant effect on dwarf systems (due to the lower escape velocity) the effect of the feedback algorithms on a Milky Way prototype is contrasted against a model of the dwarf galaxy NGC 6503. Following this investigation, a series of cosmological simulations is conducted to examine whether the conclusions from the isolated simulations hold in a cosmological environment. Particular attention is paid to the rotation curves and angular momentum transport between the galaxy and halo. The structure of the paper is as follows: In section 2, important features of the numerical technique are reviewed. In sections 3-4, the star formation prescription is presented and a detailed analysis of its performance on isolated test objects presented conducted. Results from simulations are reviewed in section 4, and a summary of the findings is given in section 6. ## 2. Numerical Method An explicit account of the numerical method, including the equation of motion and artificial viscosity used, is presented in Thacker et al. (1998). The code is a significant development of the publically available HYDRA algorithm (Couchman, Thomas & Pearce 1995). The main features of the method are summarized for clarity. Gravitational forces are evaluated using the adaptive Particle-Particle, Particle-Mesh algorithm (AP<sup>3</sup>M, Couchman (1991)). Hydrodynamic evolution is calculated using SPH. Notable features of algorithm relevant to the hierarchical simulations follow, * The neighbour smoothing is set to attempt to smooth over 52 neighbour particles, usually leading to a particle having between 30 and 80 neighbours. This number of neighbors assures a stable integration and reduces concern over the relative fluctuation in neighbor counts. * The minimum hydrodynamic resolution scale is set by $`h_{min}=ϵ/2`$ where $`ϵ`$ is the gravitational softening. If the minimum gravitational resolution is $`2ϵ`$ and the minimum hydrodynamic resolution $`4h_{min}`$ then the two of them match with this definition of $`h_{min}`$. * Once the smoothing length of a particle reaches $`h_{min}`$, all particles within $`2h_{min}`$ are smoothed over. Thus at this scale the code changes from an adaptive to nonadaptive scheme. This avoids mismatched gravitational and hydrodynamic forces at scales close to that of the resolution (Bate & Burkert (1997)). Radiative cooling is implemented using an assumed 5% $`Z_{}`$ metallicity. The precise cooling table is interpolated from Sutherland and Dopita (1993). Radiative cooling is calculated in the same fashion as discussed in Thomas & Couchman (1992). ## 3. Star Formation Prescription ### 3.1. Implementation details The star formation algorithm is based on that presented in Mihos and Hernquist (1994). Kennicutt (1998) has presented a strong argument that the Schmidt Law, with star formation index $`\alpha =1.4\pm 0.15`$, is an excellent model. It characterizes star formation over five decades of gas density, although it does exhibit significant scatter. For computational efficiency, a Lagrangian version of the Schmidt Law is used, that corresponds to a star formation index $`\alpha =1.5`$, $$\frac{dM_{}}{dt}=C_{\mathrm{𝑠𝑓𝑟}}\rho _g^{1/2}M_g,$$ (1) where $`C_{\mathrm{𝑠𝑓𝑟}}`$ is the star formation rate normalization, the $`g`$ subscripts denotes gas and the $``$ subscript stars. Assuming approximately constant volume over a time-step, both sides may divided by the volume and the standard Schmidt Law with index $`\alpha =1.5`$ is recovered. A value of $`\alpha =1.5`$ is preferred since it leads to the square root of the gas density, which is numerically efficient and the value is within the error bounds. As written the constant $`C_{\mathrm{𝑠𝑓𝑟}}`$ has units of $`\rho ^{1/2}t`$, but can be made dimensionless by multiplying by $`(4\pi G)^{1/2}`$. Hence, equation 1 may be written with dimensionless constants, leading to the following form, $$\frac{dM_{}}{dt}=\sqrt{4\pi G}c^{}\rho _g^{1/2}M_g=\frac{c^{}M_g}{t_{ff}},$$ (2) where $`c^{}`$ is the dimensionless star formation rate (Katz (1992)). The range for this parameter is reviewed in section 3.4. Limiting the star formation rate by the local cooling time-scale was not considered. The reason for this is that typically the cold dense cores which form stars are at the effective temperature minimum of the simulation for non-void regions, namely $`10^4`$ K, which corresponds to the end of the radiative cooling curve. Star formation is allowed to proceed in regions that satisfy the following criteria, 1. the gas exceeds the density limit of $`2\times 10^{26}\mathrm{g}\mathrm{cm}^3`$ 2. the flow is convergent, ($`.𝐯<0`$) 3. the gas temperature is less than $`3\times 10^4`$ K 4. the gas is partially self-gravitating, $`\rho _{gas}>0.4\rho _{dm}`$ The first criterion associates star formation with dense regions (regardless of the underlying dark matter structure). The second criterion is included to link star formation with regions that are collapsing. The third prevents star formation from occurring in regions where the average gas temperature is too high for star formation. The final criterion is particularly relevant to cosmological simulations since it limits star formation to regions where the dynamics are at least partially determined by the baryon density. Since the mass scales probed by the simulation are considerably larger than the mass scales of self-gravitating cores in which star formation would occur, we use a pre-factor of 0.4 to help compensate for this disparity. Representing the growth of the stellar component of the simulation requires compromises. It is clearly impossible to add particles with masses $`dM_{}`$ at each time-step since this would lead to many millions of small particles being added to the simulation. Alternatively, a gas particle may be viewed as having a fractional stellar mass component, i.e. the particle is a gas-star hybrid (in the terminology of Mihos & Hernquist (1994)). Thus, as gas is turned into stars, the stellar mass increases while the gas mass decreases. Mihos and Hernquist take this idea to its limit by only calculating gas forces using the gas mass of a particle (gravitational forces are unaffected because of mass conservation). A drawback of this method is that it still enforces a collisional trajectory on the collisionless stellar content of a particle. While this is a good description of real physics, since stars take a number of galactic rotations to depart from the gas cloud in which they are born, it is in strong disagreement with what would happen to collisional and collisionless particles in a simulation. To decide when to form stars, Katz (1992) and Steinmetz & Muller (1994, 1995) use a star formation efficiency. Once the star mass of the gas particle reaches a set (efficiency) percentage of the gas particle mass then a star particle is created with that mass value. This also leads to the potential spawning of very many gas particles. As a compromise, the scheme used in this work is as follows: Once $`\dot{M}_{}`$ has been evaluated, the associated mass increase over the time-step is added to the ‘star mass’ of the hybrid gas-star particle. Note, in this model the ‘star mass’ is a ‘ghost’ variable and does not affect dynamics in any way, i.e. until the star particle is created the particle is treated as entirely gaseous when calculating hydrodynamic quantities. Two star particles (of equal mass) can be created for every gas particle. This ensures that feedback events occur frequently since we do not have to wait for the entire gas content of a region to be consumed before feedback occurs, and at the same time prevents the spawning of too many star particles. The creation of a star particle occurs when the star mass of a gas-star particle reaches one half that of the mass of the initial gas particles. The gas-star particle mass is then decremented accordingly. The second star particle is created when the star mass reaches 80% of half the initial gas mass. SPH forces are calculated using the total mass of the hybrid particles. Note, in all the following sections, “gas particle” should be interpreted meaning “gas-star hybrid particle”. These assumptions yield a star formation algorithm where each gas particle has a star formation history. This assumption is motivated because cloud complexes in any galaxy have an associated star formation history. There are two notable drawbacks to this algorithm. Firstly, since star particles are only formed when the star mass exceeds a certain threshold there is a delay in forming in stars. As a consequence, prior to the star particle being formed, the SPH density used in the calculation of the SFR will be overestimated. By selecting the star mass to be one half that of the gas mass this problem is reduced, but it is not removed. Secondly, until the first star particle is spawned, the trajectory of the stellar component is entirely determined by the gas dynamics. Once a star particle is created the associated feedback must be evaluated. A simple prescription is utilized, namely that for every $`100M_{}`$ of stars formed there is one supernovae which contributes $`10^{51}`$ erg to the ISM. This value is used in Sommer-Larsen et al. (1998), and feeds back $`5\times 10^{15}`$ erg g<sup>-1</sup> of star particle to the surrounding ISM. Since this value is a specific energy, a temperature can be associated with feedback regions (see section 3.2). Navarro & White (1993), using a Salpeter IMF, with power law slope 1.5 and mass cut-offs at 0.1 and 40 $`M_{}`$, derive that $`2\times 10^{15}`$ erg g<sup>-1</sup> of star particle created is fed back to the surrounding gas. The actual value is subject to the IMF, but scaling between values can be achieved using a single parameter which we label $`e^{}`$. For $`e^{}=1`$ the energy return is $`5\times 10^{15}`$ erg g<sup>-1</sup>, while $`e^{}=0.4`$ gives the Navarro & White value. Only brief attention is paid to changing this parameter since it is more constrained than the others in the model. Variations in metallicity caused by the feedback process (Martel & Shapiro (1998)) are not considered, as this involves a complicated feedback loop involving the gas density and cooling rate. ### 3.2. Energy feedback The following sections describe each of the feedback algorithms considered. #### 3.2.1 Energy smoothing (ES) The first of the methods is comparatively standard: the total feedback energy is returned to the local gas particles. For simplicity in the following argument, it is assumed a tophat kernel is used to feedback the energy over the nearest neighbour particles. The number of neighbours is constrained to be between 30 to 80 which may divorce the feedback scale from the minimum SPH smoothing scale set by $`h_{min}`$. This occurs in the densest regions where the average interparticle spacing becomes significantly less than $`h_{min}`$. This actually renders feedback more effective in these regions than it would be if all the energy were returned to the particles with in $`2h_{min}`$ region, but the effect is not significant. Given these assumptions for star formation, and working in units of internal energy, it is possible to evaluate the temperature increase, $`\mathrm{\Delta }T`$ of a feedback region as follows, $$\mathrm{\Delta }T=\frac{2}{3}\frac{\mu m_p}{k}\frac{E_{SN}M_{}}{N_sM_g},$$ (3) which given $`N_s=52`$, $`E_{SN}=5\times 10^{15}\mathrm{erg}\mathrm{g}^1`$, and $`M_{}/M_g=1/2`$ yields, $$\mathrm{\Delta }T2.4\times 10^7\frac{M_{}}{N_sM_g}2.3\times 10^5\mathrm{K}.$$ (4) Clearly this boost may be increased or decreased by altering any one of the variables $`E_{SN}`$, $`N_s`$ and the ratio $`M_{}/M_g`$. Keeping $`E_{SN}`$ constant, but reducing $`N_s`$ to 32 and increasing $`M_{}/M_g`$ to 1 would yield $`\mathrm{\Delta }T7\times 10^5\mathrm{K}`$, demonstrating the sensitivity of feedback to the SPH smoothing scale. #### 3.2.2 Single particle feedback (SP) Alternatively, all of the energy may be returned to a single SPH particle. Gerritsen (1997) has shown that this is an effective prescription in simulations of evolved galaxies, yielding accurate morphology and physical parameters. In this case the temperature boost is trivially seen to be $$\mathrm{\Delta }T2.4\times 10^7\mathrm{K}.$$ (5) as $`N_s=1`$ and $`M_{}/M_g=1`$. There is one minor problem in that when the gas supply is exhausted, the mechanism has no way of returning the energy (unless one continues to make star particles of smaller and smaller masses). As a compromise the nearest SPH particle is found and the energy is given to this particle. #### 3.2.3 Temperature smoothing (TS) The final feedback mechanism considered is one that accounts for the fast radiative losses by increasing the energy input. The first step is to calculate the temperature a single particle would have if all the energy were returned to it (as in the SP model). Then this value is smoothed over the local particles using the SPH kernel. This method leads to a vastly higher energy input than the others and represents a case of extreme feedback (essentially 52 times the ES feedback). For the isolated simulations, the cooling mechanism (see below) was not adjusted in this model since the feedback regions in the disk have cooling time of several time-steps (at least 10). In the cosmological simulations, the alternative cooling mechanisms were considered since the cooling time is only a few time-steps (Katz (1992); Sommer-Larsen, Gelato & Vedel (1998)). #### 3.2.4 Preventing immediate radiative energy losses As was previously discussed, Katz (1992) was the first to show that feedback energy returned to the ISM is radiated away extremely quickly in high density regions. This is a result of the characterisitic dynamical time of the simulation being far longer than that of cooling. The first method for preventing the immediate radiative loss of the feedback energy is to alter the density value used in the radiative cooling mechanism. This change is motivated by Gerritsen’s (1997) tests on turning off radiative cooling in regions undergoing feedback. Assuming pressure equilibrium between the ISM phases (which is not true after a SN shell explodes but is a good starting point) one may derive the estimated density that the region would have after the SN shell has exploded. If the local gas energy is increased by $`E_{SN}`$, then the perfect gas equation of state yields $$\rho _{est}=\frac{E_i\rho _i}{E_i+E_{SN}}.$$ (6) Following a feedback event the estimated density is allowed to decay back to it’s local SPH value with a half-life $`t_{1/2}`$. Fig. 1.—Evolution of the cooling density and the SPH density following a single feedback event in the ESna scheme. $`t_{1/2}=5`$ Myr, and clearly the values converge within $`2t_{1/2}`$. Note the small initial drop in the SPH density in response to the feedback energy, followed by a slower expansion. The initial cooling density is approximately 5% of the SPH density, consistent with the feedback temperature of $`2\times 10^5`$ K in an ambient 10,000 K region. To calculate the decay rate (i.e. to predict the cooling density at the time-step $`n+1`$ from that at time-step $`n`$) the following function is used $$\rho _{cool}^{n+1}=\{\begin{array}{cc}\rho _{cool}^n+dt\times \mathrm{\Delta }\rho ^n\left(t_f^2/t_{1/2}^3\right)e^{0.33\left(t_f/t_{1/2}\right)^3},\hfill & t_ft_{1/2}\text{;}\hfill \\ \multicolumn{2}{c}{}\\ \rho _{SPH}^n\mathrm{\Delta }\rho ^ne^{0.693dt/t_{1/2}},\hfill & t_{1/2}t_f<3t_{1/2}\text{;}\hfill \\ \multicolumn{2}{c}{}\\ \rho _{SPH}^{n+1},\hfill & 3t_{1/2}t_f\text{ ,}\hfill \end{array}$$ (7) where $`t_f`$ is time since the feedback event occurred, $`\mathrm{\Delta }\rho ^n=\rho _{SPH}^n\rho _{cool}^n`$, and $`dt`$ is the time-step increment. Once a region passes beyond $`3t_{1/2}`$, the cooling density is forced to be equivalent to the SPH density, although usually the values converge within $`2t_{1/2}`$. This comparatively complex function was chosen because in a simple exponential decay model, the cooling density increases by the largest amount immediately following the feedback event. To have any effect the cooling density must be allowed to persist at its low value for a reasonable period of time. In figure 1 the two densities calculated after a single feedback event are compared. This cooling mechanism is denoted by a suffix na (for non-adiabatic) on the energy input acronym. Springel and White (in prep) have considered a similar model (Pearce, private communication). Gerritsen allowed his SPH particles to remain adiabatic for $`3\times 10^7`$ years, approximately the lifetime of a $`8M_{}`$ star. It is difficult to argue what this value should be and hence the $`t_{1/2}`$ parameter space was explored. Note that if during the decay period another feedback event occurs in the local region, the density value used is the minimum of the current decaying one and the new calculated value. As a second method for preventing radiative losses Gerritsen’s idea was utilized: make the feedback region adiabatic. This is easily achieved in our code by using the same mechanism that calculates the estimated density value. Provided the estimated density value is less than half that of the local SPH value then the particle is treated as adiabatic. Above this value the estimated density is used in the radiative cooling equation. This cooling mechanism is denoted with a suffix a. ### 3.3. Comparison of methods To test each of the feedback methods and gain insight into their effect on the local ISM, a single feedback event was set up within a prototype isolated Milky Way galaxy. The evolution of the particles within $`3h`$ of the feedback event were followed. The time evolution of the temperature versus radius for each scheme is shown in figure 2. #### 3.3.1 Qualitative discussion The adjusted cooling mechanism has little effect on the ES run because the estimated density is not low enough to increase the cooling time significantly beyond the length of the time-step. Including a prefactor of 0.1 in the equation, so that the estimated density is significantly lower, does increase the cooling time sufficiently. However, since introducing the adiabatic phase allows the feedback energy to induce expansion, we prefer this method, rather than trying to adjust the estimated density method. For the SP run the density reduction is much higher (since the total energy, $`E_{SN}`$, is applied to the single particle) and hence the mechanism does allow the particle to remain hot. There is little perceptible difference between the SPa and SPna profiles. Both SP feedback, TS, and ESa induce noticeable expansion of the feedback region. Thus after the heat input has been radiated away, the continued expansion introduces adiabatic cooling (since the temperature of the region falls below the $`10,000`$ K cut off of the cooling curve). It is particular noticeable in the TS plot where the low temperature plateau continues to widen quite drastically and this is manifest in the simulation with the appearance of a large bubble in the disk. Caution should be exercised in interpreting these bubbles in any physical manner, since their size is set solely by the resolution scale of the SPH. ESa also produces bubbles, but due to the lower temperature there is less expansion. Single particle feedback produces the smallest bubbles and often ejects the hot particle vertically from the disk. This occurs with regularity since the smoothing scale is typically larger than the disk thickness: if the hot particle resides close to the edge of the disk the pressure forces from the surrounding particles will be asymmetric resulting in a strong ‘push’ out of the disk. Note that this is phenomenologically similar to the mechanism by which SN gas is ejected from disks (McKee & Ostriker (1977)), although this should not be over-interpreted. The only scheme which stands out in this investigation is ESna: the cooling mechanism fails to prevent drastic radiative losses. The remainder of the algorithms produce an effect on both the thermal properties of the ISM and its physical distribution. #### 3.3.2 Effect of methods on time-step criterion Since our code does not have multiple time-steps, it is important to discern whether one method allows longer time-steps than another. This is a desirable feature since it reduces the wall-clock time for simulations. Of course an algorithm which has a fast wall-clock time but produces poor results would never be chosen, however for two algorithms with similar results this criterion provides a useful parameter to choose one over the other. A comparison of SPna, TS, ESna and the no feedback (NF) run is displayed in figure 3. The simulation time versus the number of time-steps was compared for data from the Milky Way prototype runs (see section 4.1). The SP methods produce the shortest time-step, requiring almost double the number of time-steps than the NF model. The ES variants require only 10% more time-steps than the run without feedback. In SP feedback the acceleration felt by the hot particle limits the time-step significantly. TS also requires more time-steps than ES due to the rapid expansion of feedback regions. Typically though, the number of time-steps required is somewhat less (20%) than that for single particle feedback. ### 3.4. Explored parameter space of the algorithm All models exhibit dependencies on the free parameters $`c^{}`$ and $`e^{}`$, corresponding to the SFR normalization and efficiency of the feedback energy return. The models which use a modified cooling formalism also exhibit a dependence upon $`t_{1/2}`$, the approximate half-life of feedback regions. To simplify matters, an ensemble with $`e^{}=0.4`$ (the value used in Navarro & White 1993) was run, allowing us to concentrate on the effect of the $`c^{}`$ and $`t_{1/2}`$. To determine the effect of varying $`e^{}`$, two more simulations with $`e^{}=1`$ were run. | Run | $`c^{}`$ | $`t_{1/2}^\mathrm{a}`$ | $`e^{}`$ | $`N_{step}^\mathrm{b}`$ | | --- | --- | --- | --- | --- | | 5001 | 0.001 | 0 | 0.4 | 1999 | | 5002 | 0.003 | 0 | 0.4 | 2001 | | 5003 | 0.01 | 0 | 0.4 | 1924 | | 5004 | 0.03 | 0 | 0.4 | 1977 | | 5005 | 0.1 | 0 | 0.4 | 1989 | | 5006 | 0.3 | 0 | 0.4 | 2140 | | 5007 | 1.0 | 0 | 0.4 | 2087 | | 7001 | 0.033 | 1. | 0.4 | 1989 | | 7002 | 0.033 | 5. | 0.4 | 2000 | | 7003 | 0.033 | 10. | 1.0 | 1947 | | 7004 | 0.033 | 1. | 1.0 | 1938 | | 7005 | 0.033 | 10. | 0.4 | 1940 | Table 1: Summary of star formation parameter space simulations. The simulations were of a rotating cloud collapse (see Thacker et al (1998)). <sup>a</sup> $`t_{1/2}=0`$ denotes that feedback was removed from the simulation. <sup>b</sup> The number of steps are given to t=1.13, the final point of the parameter space plot. #### 3.4.1 Simple collapse test To gain an understanding of the algorithms in a simple collapse model (that also may be run in a short wall-clock time) the rotating cloud collapse model of Navarro and White (1993, also see Thacker et al (1998)) was utilized. Such models actually bear little resemblance to the hierarchical formation picture, but they do allow a fast exploration of the parameter space. For this test the self-gravity requirement was removed. The reason for this is that in cosmological simulations it is virtually guaranteed that the gas in a compact disk will be self-gravitating. This is due to the low number of dark matter particles in the core of the halo relative to the number of gas particles. The most important parameter in the star formation model is the $`c^{}`$ parameter since it governs the SFR normalization. Therefore, an ensemble of models was run with $`c^{}[0.001,1]`$ (and $`e^{}=0.4`$). The secondary parameter in the model, $`t_{1/2}`$, is expected to have comparatively little effect on the star formation rate (due to the low volume factor of regions undergoing feedback). Hence only a range of plausible alternatives were considered, namely $`t_{1/2}=1,5,10`$ Myr, in the ESa model. The simulation parameters are detailed in table 1. #### 3.4.2 Results Figure 4 displays a plot of the $`c^{}`$ parameter space, showing SFR and $`c^{}`$ versus time. Feedback was effectively turned off in this simulation by reducing the energy return efficiency to $`10^7`$. Although a severe amount of smoothing has had to be applied (a running average over 40 time-steps, followed by linear interpolation on to the grid) there are a number of interesting results. The time at which the peak SFR occurs is almost constant across all values of $`c^{}`$. This is an encouraging result since it indicates that the time at which star formation peaks is dictated by dynamics and not by the parameters of the model (at least without feedback). In fact the SFR peak time corresponds to the time when the collapse model reaches its highest density, following this moment a significant amount of relaxation occurs and the gas has a Fig. 3.—Effect of feedback scheme on the time-step selection in the simulation. Data from the Milky Way prototype runs is shown. Twice as many time-steps are required for SP feedback compared to runs without feedback. The ESa and SPa runs are not shown, but are approximately 5% lower than the respective runs without the adiabatic period. Fig. 4.—Dependence of the SFR on the $`c^{}`$ parameter in a model with no feedback. The data for seven runs was linearly interpolated to form the plotted surface. The time of the peak SFR moves very little with changing $`c^{}`$, and almost all the models can be fitted to exponential decay models following the peak SFR epoch. lower average density. Note, however, that this is an idealized model with no feedback and a uniform collapse. Figure 5 shows the dependence of the SFR on the $`t_{1/2}`$ and $`e^{}`$ parameters. To detect trends in the SFR, a running average is shown, calculated over 40 time-steps. Clearly, there is little distinction between the runs with $`t_{1/2}=1`$ and 10. This can be attributed to the low volume fraction of regions undergoing feedback. It is interesting to note that the SFR is more sensitive to the amount of energy returned, dictated by the $`e^{}`$ parameter, than the lifetime for which this energy is allowed to persist. The line corresponding to $`e^{}=1`$ (the standard energy return value of $`5\times 10^{15}\mathrm{erg}\mathrm{g}^1`$) does not have the secondary and tertiary peaks in the SFR exhibited by the $`e^{}=0.4`$ runs. This is probably attributable to the $`.𝐯`$ criterion: the $`e^{}=1`$ Fig. 5.—Dependence of the SFR on the $`t_{1/2}`$ and $`e^{}`$ parameters (for the ESa algorithm). Comparing the runs with the same $`e^{}`$ parameter shows that varying $`t_{1/2}`$ from 1 to 10 has no noticeable effect. Conversely, changing $`e^{}`$ from 0.4 to 1.0 removes the secondary and tertiary peaks in the SFR. run produces enough heating to provide a significant amount of expansion rendering $`.𝐯0`$ for the first feedback region. Note that since less of the gas is used at early times, the SFR at later epochs is higher. In summary, while the $`c^{}`$ parameter clearly sets the SFR normalization, it does not change the epoch of peak star formation. Further, the $`t_{1/2}`$ parameter has little effect on the overall SFR due to the low volume fraction of feedback regions in the evolved system. ## 4. Application to isolated ‘realistic’ models This section reports the results of applying the algorithm to idealized models of mature isolated galaxies. These models are created to fit the observed parameters of such systems, i.e. the rotation curve and disk scale length. The characteristics of each model are discussed within the sections devoted to them. In this investigation the relative gas to stellar fraction is low, compared to the primordial ratio, and thus star formation is not as rapid as would be expected in the early stages of the cosmological simulations (section 5). Both models were supplied by Dr. Fabio Governato. We note that they both have a sufficiently high particle number ($`10^4`$ SPH particles) to represent the gas dynamic forces with reasonable accuracy (Steinmetz & Muller (1993); Thacker et al (1998)). A summary of the simulations is presented in table 2. ### 4.1. Milky Way prototype The first prototype model is an idealized one of the Galaxy. It is desired that the model should reproduce the measured SFR $``$ 1 M yr<sup>-1</sup> and also the velocity dispersion in the disk. Evolved galaxies have a lower relative gas content than protogalaxies. Further, because of hierarchical clustering, they are significantly more massive. Hence feedback is expected to have less effect on this model than on protogalaxies formed in simulations of hierarchical merging. #### 4.1.1 Model Parameters The Milky Way prototype contains stars, gas and dark matter. The total mass of each sector is $`5\times 10^{10}`$ M, | Run | Simulation object<sup>a</sup> | feedback<sup>b</sup> | $`N_{step}^\mathrm{c}`$ | $`N_{SPH}`$ | | --- | --- | --- | --- | --- | | 1001 | NGC 6503 | none | 3010 | 10240 | | 1002 | NGC 6503 | SPna | 3453 | 10240 | | 1003 | NGC 6503 | SPa | 3986 | 10240 | | 1004 | NGC 6503 | TS | 5345 | 10240 | | 1005 | NGC 6503 | ESna | 3453 | 10240 | | 1006 | NGC 6503 | ESa | 4535 | 10240 | | 2001 | Milky Way | none | 387 | 10240 | | 2002 | Milky Way | SPna | 952 | 10240 | | 2003 | Milky Way | SPa | 943 | 10240 | | 2004 | Milky Way | TS | 723 | 10240 | | 2005 | Milky Way | ESna | 424 | 10240 | | 2006 | Milky Way | ESa | 453 | 10240 | | 6001 | Cosmological | SPa | 4792 | 17165 | | 6002 | Cosmological | SPna | 4710 | 17165 | | 6003 | Cosmological | ESa | 4319 | 17165 | | 6004 | Cosmological | ESna | 4319 | 17165 | | 6005 | Cosmological | TSna | 4335 | 17165 | | 6006 | Cosmological | TSa | 4322 | 17165 | | 6007 | Cosmological | NF | 4341 | 17165 | | 6008 | Cosmological | NSF | 4314 | 17165 | | 6010 | Cosmological | TS | 4342 | 17165 | Table 2: Summary of the main simulations using realistic models. <sup>a</sup>‘Cosmological’ refers to the object formed being derived from a cosmological simulation. <sup>b</sup>SPa=Single particle adiabatic period, SPna=single particle no adiabatic period but adjusted cooling density, ESna=Total energy smoothing with adjusted cooling density but no adiabatic period, ESa=Total energy smoothing with adiabatic period, TS=Temperature smoothing (normal cooling), NF=no feedback, NSF=no star formation. <sup>c</sup>For the cosmological simulations the number of time-steps to $`z=1.0`$ is given. Since the isolated simulations were not run to the same final time, the average number of time-steps per 100 Myr is shown. $`9\times 10^9`$ M, and $`3\times 10^{11}`$ M respectively. 11980 particles were used to represent the stars, 10240 to represent the gas and 10240 to represent the stars. The individual particle masses were $`4\times 10^6`$ M, $`9\times 10^5`$ M, and $`3\times 10^7`$ M respectively. The (stellar) radial scale length was 3.5 kpc and the scale height 0.6 kpc. Density and velocities were assigned using the method described in Hernquist (1993). The maximal radius of the dark matter halo was 85 kpc. The artificial viscosity was not shear-corrected in this simulation since the simulation was integrated through only slightly more than two rotations and the particle resolution in the object of interest is quite high. A comparatively large softening length of 0.5 kpc was used, rendering the vertical structure of the disk poorly resolved. However, this is in line with the softening lengths that are typically used in cosmological simulations (of order 2 kpc). Shorter softening lengths allow higher densities in the SPH, which in turn leads to higher SFRs. The self-gravity requirement was again removed. #### 4.1.2 Results In figure 6, the gas particle distributions are shown for the NF, SPa, ESa and TS runs. Of the versions not shown, ESna has a smooth disk since the feedback regions do not persist as long and the SPna disk resembles that from the SPa run (see section 3.3.1). TS produces the most significant disturbance in the disk, which is to be expected given that it injects more energy into the ISM than the other methods. For TS feedback, 7% of the disk gas had been ejected (falls outside an arbitrary horizontal 6 kpc band) by t=323 Myr rising to 14% by t=506 Myr. Note that the amount of ejected gas is calculated relative to the total gas in the simulation at the time of measurement. This is a decreasing function of time, but is similar for all simulations since the SFRs differ little. Particle ejection velocities, $`v_z`$, were close to 300 km s<sup>-1</sup>, although some did achieve escape velocity ($`500`$ km s<sup>-1</sup> at solar radius). Hence, while TS can project particles out of the halo (‘blow-away’) it preferentially leaves them bound in the halo (‘blow-out’). SP feedback (both SPa and SPna) has a similar evolution, but only tends to eject the single heated particle during each feedback event, thus leading to a lower mass-loss rate (1% of the disk gas had been ejected by t=323 Myr for both versions). Particles were often ejected with $`v_z>600`$ km s<sup>-1</sup>, which is larger than the escape velocity, leading to a proportionally stronger tendency for blow-away. ESa also ejects particles from the disk (0.4% ejected by t=323 Myr), although in general the particles have lower velocities ( $`v_z200`$ km s<sup>-1</sup>) than the particles ejected by either TS or SP. Hence, almost all of the ejected gas remains bound to the system, i.e. ES leads only to blow-out. ESna does not eject particles since the feedback regions cool sufficiently fast (0.01% ejected by t=323 Myr). The SFRs for three of the simulations (NF,TS and ESa) are plotted in figure 7. Most noticeable is the reduction in the SFR produced by TS and ESa (TS is 35% lower than no feedback at t=500 Myr, ESa is 10% lower). This is due to three factors; (a) the ejection of matter from the disk depletes the cold gas available for star formation (see figure 6), (b) smoothing feedback energy leads to spatially extended ‘puffy’ feedback regions in the disk, which in turn leads to a lower average density, and hence lower SFR, (c) particles in the feedback regions will typically be above the temperature threshold, which prevents star formation which further reduces the SFR. For single particle feedback the lower mass loss rate leads to a higher SFR than for the TS or ESa runs. Of the versions not plotted, ESna resembles the no feedback run (SFR approximately 3-4% lower on average), since most of the energy is rapidly radiated away. SPna resembles the SPa since the feedback events produce very similar effects (see section 3.3.1). To calculate radial profiles of the disks, an arbitrary plane of thickness 6 kpc was centered on the disk. This thickness ensured that the stellar bulge was contained within the band. Radial binning was then performed on this data set using cylindrical bins. In figure 8, gas rotation curves are compared for the simulations at t=323 Myr. To provide a fairly accurate depiction of the rotation curve that would be measured the rotation curve was calculated by radial averaging $`|𝐫\times 𝐯|/|𝐫|`$ rather than by calculating the circular velocity from $`\sqrt{GM(<R)/R}`$. The main drawback of this method is that in the core regions, where there are few particles in the bins, the measurement can become ‘noisy’. Clearly, figure 8 shows that there is little difference between schemes (a maximum of 9% at 4 scale lengthsignoring the Fig. 7.—SFRs for the Milky Way prototype (time-averaged over 160 time-steps to show trend). The SP variants are not shown since their evolution is similar to that of the ES version plotted. TS produces a significant (35% at t=500 Myr) reduction in the SFR as compared to no feedback. ES also reduces the SFR, but has a less significant effect (10% reduction at t=500 Myr versus no feedback) than TS. under-sampled central values). At large radii the curves match precisely since there are no feedback events in the low-density outer regions of the disk, except for the TS run where a feedback event has ejected particles to the outer regions. Comparing to the initial rotation curve (not shown), the disk has clearly relaxed, extending both in the tail and toward the center. The curves were also examined at t=507 Myr (for those simulations integrated that long) and similar results were found with maximum differences being in the 10% range. A more telling characteristic is the gas radial velocity dispersion. Unfortunately it is difficult to relate the measurements made here to those of molecular clouds (Malhotra (1995)), primarily because the mass scales are significantly different. Nonetheless, it is interesting to compare each of the separate feedback schemes. In figure 8 the radial velocity dispersion, $`\sigma _r`$, is plotted for three of the simulations at t=323 Myr (again only considering matter within the 6 kpc band). Temperature smoothing produces a large amount of dispersion due to the excessive energy input (interior to 8 scale lengths it varies between being 40% to 300% higher than other values). Note that there is a direct correlation between a higher velocity dispersion and a lowered measured rotation curve. This is asymmetric drift: feedback events produce velocity dispersion which in turn increases the relative drift speed, $`v_a`$, between the gas and the local circular velocity (Binney & Tremaine (1987)). The remaining algorithms (SPa, SPna, ESa, ESna) exhibit similar velocity dispersions. Thus, for all but the TS algorithm, the bulk dynamics remain the most important factor in determining the velocity dispersion. Although the velocity dispersion presented here is not directly compatible with that of the value for local molecular clouds, it is interesting to note that measurements for the Milky Way suggest $`\sigma _{v_{cloud}}=9\pm 1`$ km s<sup>-1</sup>(Malhotra (1995)). Due to the finite size of the computational grid, the code was unable to follow all of the simulations to the desired final epoch (500 Myr). This limitation was most noticeable in the SP simulations where ejected gas particles reached the edge of the computational domain within 320 Myr. It is possible to remove these particles from the simulation since they are comparatively unimportant to the remainder of the simulation. However maintaining an accurate integration was considered to be more important. Fig. 8.—Rotation curves and radial velocity dispersions for the Milky Way prototype at t=323 Myr, for the NF, SPa, ESa and TS runs. There is only a marginal difference between rotation curves because the radial averaging smooths out the effect of inhomogeneous feedback regions. The TS gas exhibits a 8% reduction in the rotation curve at a radius of 4.5 scale lengths due to asymmetric drift. The ESa rotation curve is reduced by only 4% at this radius. The TS algorithm exhibits significantly higher velocity dispersion than any of the other variants. This is attributable to the large energy input driving winds that strongly affect the disk structure (visible in figure 6 as large holes in the disk). All the other algorithms differ only very marginally. #### 4.1.3 Summary Temperature smoothing (TS) is clearly the most violent feedback method, producing an SFR lower than the other algorithms and also tending to evaporate the disk. Given the comparatively large escape velocity of the Milky Way (lower limit of 500 km s<sup>-1</sup> at the solar radius), this is an unrealistic model because such dramatic losses are expected only in dwarf systems. Of the remaining algorithms, the single particle (SP) versions produce reasonable physical characteristics, but have the disadvantage of requiring a large number of time-steps. The energy smoothing variant with an adiabatic period (ESa) appears to be the best compromise in these simulations. It does not require an excessive number of time-steps while the disk morphology and evolution are within reasonable bounds: there is no excessive blow-out or blow-away. ### 4.2. Dwarf prototype The second model is an idealized version of NGC 6503. Dwarf systems are expected to be more sensitive to feedback due to their low mass, and consequently lower escape velocity. In simulations, the over-cooling problem suggests that to form a large disk system from the merger of small dwarfs, the dwarf systems must have significant extent (ideally similar to that for an adiabatic system, Weil, Eke & Efstathiou (1998)). Feedback is currently believed to be the best method for achieving this. Given that for NGC 6503 $`v_c110`$ km s<sup>-1</sup> a lower bound on the escape velocity of the system is 155 km s<sup>-1</sup>. Detailed numerical studies of NGC 6503 have been conducted by Bottema and Gerritsen (1997) and Gerritsen (1997). The motivation in this investigation is different to the previous ones which attempted to explain the observed dynamics of NGC 6503. In contrast, this investigation attempts to determine bulk properties at comparatively low resolution, in accordance with that found in cosmological simulations. #### 4.2.1 Model Parameters As for the Milky Way prototype, this model contains stars, gas and dark matter, with the total masses of $`3\times 10^9`$ M, $`1\times 10^9`$ M, and $`5\times 10^{10}`$ M respectively. 10240 particles were used in each sector, yielding individual particle masses of $`3\times 10^5`$ M, $`1\times 10^5`$ M, and $`5\times 10^6`$ M, respectively. The radial scale length of the simulation was 1.16 kpc, and the scale height 0.1 kpc. Gas density and particle velocities were assigned in the same fashion as the Milky Way prototype. The artificial viscosity was not shear-corrected for the same reasons as the Milky Way protoype. Six simulations were run, each using a different method of feedback (including no feedback as the control experiment). The method used in each simulation and the number of time-steps per 100 Myr are summarized in table 2. #### 4.2.2 Results As in the Milky Way simulations, the SP feedback models produced significant blow-away at the outset of the simulation and particles ejected from the disk rapidly escaped from the halo. Consequently, the evolution of these systems had to be halted at very early times (close to 200 Myr). Of the remaining algorithms, only TS was not integrated to at least 500 Myr. Thus, the following analysis concentrates on the ES variants. Figure 9 shows the distribution of gas particles in the x- and z-projections and also the z-distribution measured vertically in bins, at t=230 Myr (note that although sufficient time has elapsed for feedback events to occur the disks still exhibit virtually identical rotation curves). Due to the comparatively long softening (300 kpc) length used in the simulation, there is a significant amount of relaxation from the initially ‘thin’ distribution, which is approximately 250 pc wide. Since in the simulation code the SPH resolution is at least twice the gravitational softening length, the disk was expected to fatten to 600 pc. Figure 9 shows that this is observed (in the run with no feedback). Once feedback is included, and matter is ejected from the disk, the z-distributions develop populated tails due to particles orbiting high in the potential well. The most severe example of this being the TS variant, which rapidly ejects particles leading to significant mass loss from the disk (both blow-away and blow-out occur). The TS algorithm produces extremely large bubbles in the disk. One bubble had a radius of almost 0.7 kpc, which is 60% of the disk scale length, while for the Milky Way prototype larger bubbles had a radius of about 0.8 kpc, approximately 40% of the scale length. The absolute size of these bubbles relative to the disk is set by the SPH smoothing scale and hence should not be overinterpreted. However, the comparison of the dwarf versus the Milky Way model is valid since the particle resolution is approximately the same for both models. A comparison of the gas distributions for the ESa runs (dwarf vs. Milky Way) at 500 Myr shows that while in the dwarf system the gas density puffs up beyond the stellar component, it does not do this for the Milky Way prototype. As in the Milky Way runs, ES preferentially leads to blow-out, although by 500 Myr some particles were close to escaping the halo. These results show that feedback has a more significant effect on the dwarf system. To calculate radial profiles of the disks a plane of thickness 2 kpc was used, centered on the disk. As for the Milky Way prototype, the thickness was chosen to ensure that the stellar content was included within the band. The data were again binned using concentric cylinders. At t=580 Myr the gas rotation curves for ESna, ESa and no feedback remain very similar (figure 10). Both of the curves exhibit asymmetric drift relative to the run with no feedback. The maximum difference (external to a radius of one scale length) occurs at 1.6 scale lengths where the adiabatic variant has a rotation curve that is 10% lower than the no feedback run. At this radius the non-adiabatic run is only 3% lower than the run with no feedback. The outer edges of the distributions remain identical due to there being no feedback events in the low-density gas. The gas radial velocity dispersion plot (figure 10) shows that for this system ESa clearly introduces more dispersion (30% higher at a radius of 1.6 scale lengths). This is as expected: the combination of a lower escape velocity and comparatively long persistence of the feedback regions in the adiabatic variant allow the gas to escape to higher regions of the potential well. Additionally, bubble expansion in the plane of the disk persists for longer in the adiabatic variant. Notably, the run with no feedback shows an increasing velocity dispersion with radius. This can be attributed to the large softening length used, which in turn causes the SPH to smooth over a very large number of particles in the central regions (in excess of $`10N_{smooth}`$). Thus, in this region the gas distribution is dynamically cold. Figure 11 shows the time-averaged SFRs for the three runs. By t=580 Myr the non-adiabatic run had ejected 1% of its mass from the disk (relative to the remaining gas), whilst the adiabatic run had ejected 2%. Examination of the raw data shows that the strongest bursting is actually found in the no feedback run. The feedback in the other two runs keeps the disk more stable against local collapse. Of the data not plotted, TS was similar to the ESa run, and had an SFR approximately 5% lower. By t=240 Myr, 6% of the disk had been evaporated. Neither of the SP runs was integrated far enough to detect Fig. 10.—Comparison of rotation curves and radial velocity dispersions for the NGC 6503 dwarf simulation at t=580 Myr. There is little difference among all rotation curves, except in the nuclear region where feedback is more prevalent. Asymmetric drift is visible in the gas, with the ESa run having a lower rotation curve due to its higher velocity dispersion. Fig. 11.—SFRs for the dwarf prototype. The ES and no feedback runs are shown since the remainder were not integrated for a sufficient time for conclusions to be drawn. The NF, ESa, and ESna runs were time-averaged over 320 time-steps to elucidate the trend in the SFR and this is indicated by the bar symbol in the legend. Clearly the ESa run produces the lowest SFR, being approximately 20% lower than the non-adiabatic run. significant trends. #### 4.2.3 Summary Although it was not possible to integrate all the models to the desired final epoch, it was still clearly demonstrated that feedback does have a more significant impact on the dwarf system. This is evident both in the morphology (larger relative bubbles as compared the Milky Way prototype) and radial characteristics (the radial velocity dispersion is far higher relative to the no feedback run). This increased sensitivity also allows differentiation between the adiabatic and non-adiabatic methods, which was at times difficult in the Milky Way prototype. Whether these conclusions can carry over to cosmological simulation is addressed in the next section. ## 5. Cosmological simulations As discussed in the introduction, there are a number of problems that plague cosmological simulations of galaxy formation. This section examines the conjecture that, following an initial burst of star formation, feedback should be able to (1) prevent the overcooling catastrophe by suppressing massive early star formation and (2) prevent the angular momentum catastrophe, thereby allowing the formation of disks with specific angular momenta in agreement with observations. We study all of the feedback algorithms analysed in the previous sections and also include two new versions derived from combinations of the previously studied algorithms. Note that the SPH resolution ($`2\times 10^3`$ particles) in the galaxies formed in the following simulations is insufficient to resolve shocks adequately. The purpose of the simulations is to explore the parameter space and not make precise predictions about resulting galaxies. A higher resolution simulation, meeting the accuracy criteria outlined in Steinmetz & Muller (1993), will be presented in a subsequent paper. ### 5.1. Initial conditions In the SPH method, shocks are captured using an artificial viscosity. The artificial viscosity is turned on or off by the value of the $`𝐫.𝐯`$ product between each pair of particles. The angular part of this product takes a maximal value if r and v are aligned, as is the case in collapse along a Cartesian grid. Collapse along a direction not aligned along the grid leads to scatter in the $`𝐫.𝐯`$ dot product, and hence less shocking. Consequently, we believe that it is advantageous to use a set of initial conditions must be used that contains no preferred direction. ‘Glass-like’ initial conditions are used in this study. Given a hierarchical clustering scenario the first objects to form will have hundreds (at most) particles in them. Hence it is only necessary to create initial conditions which have no preferred direction on scales of the order 500 particles. The merging of the first generation progenitors occurs over scales significantly larger than a grid spacing, thus removing concerns about a preferred collapse direction for these objects. It thus makes sense to create a small periodic glass with very low noise and then tile this within the simulation box. To create the glass ‘tile’, 512 particles are placed in a periodic box. These particles are forced to be anti-correlated by not allowing any two particles to be closer than $`0.9\times N^{1/3}`$, which is 0.9 times the average inter-particle spacing. This initial condition has a very low noise level. The noise level is further reduced by evolving the glass in a (periodic) gravity-only simulation with the sign of the velocity update reversed. With this modification, the particles repel one another, and eventually relax Fig. 12.—Layering of cosmological initial conditions. Starting clockwise from the upper left panel, the top level configuration of $`32^3`$ particles is repeatedly cut and shrunk to create a hierarchy four levels deep. Gas particles are included in the central region only. to a state in which the (repulsive) potential energy is reduced to a minimum. Once the tile is fully evolved, it is replicated a number of times to form the main simulation cube. This configuration does not have any noise on scales larger than the size of tile and thus constitutes an excellent initial configuration. Because long-range tidal forces have a significant effect on the evolution of galaxies (Kofman & Pogosyan (1995)) they must be included in simulations. Unfortunately, a fixed resolution periodic box with equal number of dark matter and gas particles would require a prohibitively large number of particles. Hence we used the multiple mass technique (Porter (1985)) to overcome this problem. The hierarchical layers are constructed by successively cutting out a region of the simulation cube and replacing it with a copy of the top-level ‘grid’ cut and shrunk to the appropriate size. The first layer, for example, is constructed by removing a sphere of radius $`1/4`$ the box size and then filling that region with a sphere cut from the main simulation cube and shrunk to size. The next layer is formed by cutting a sphere of radius $`1/8`$ the box and replacing this with a similarly cut and shrunk sphere from the main simulation cube. Unfortunately this process does introduce some noise at the boundary of each region. It was thus assured that in the highest resolution region, objects of interest form sufficiently far away from the boundary with the next region. The layering process continues through four layers. To maintain mass resolution, the particles in each layer have mass $`1/8`$ that of the previous layer. Thus the mass resolution of this region is 512 times higher than the lowest resolution region, and the spatial resolution is eight times higher. Figure 12 shows the layering in detail. If a grid of $`32^3`$ particles is initially used to perform the layering, the resulting system has 77,813 dark matter particles and 17,165 SPH particles. In the high resolution region the effective resolution is $`2\times 256^3`$. Assigning the perturbations associated with the initial power spectrum is more difficult for multiple mass simulations. The particle Nyquist frequency is not constant across the simulation. If the box is loaded with modes up to Nyquist frequency of the highest resolution region, then aliasing of the extra modes will occur in the low resolution region. Hence, to prevent aliasing, the lower resolution regions must have their displacements evaluated from a force grid that is calculated using only modes up to the local particle Nyquist. Thus all of the box modes are calculated, and then modes are progressively removed by applying a top-hat filter in k-space. ### 5.2. Simulation Parameters To assign adiabatic gravitational perturbations, the linear CDM power spectrum of Bond and Efstathiou (1984) was utilized, $$P(k)=\frac{Aq}{[1+(23.1q+(11.4q)^{3/2}+(6.5q)^2)^{5/4}]^{8/5}},$$ (8) where, $$q=k/(\mathrm{\Gamma }h).$$ (9) Given a baryon fraction of 10% the shape parameter, $`\mathrm{\Gamma }`$, was calculated from Vianna and Liddle (1996) yielding $`\mathrm{\Gamma }0.41`$. The Hubble constant was set at 50 km s<sup>-1</sup> Mpc<sup>-1</sup>, yielding in $`h=0.5`$ in the standard H<sub>0</sub>=$`100h`$ km s<sup>-1</sup> Mpc<sup>-1</sup> units. The normalization constant, $`A`$, was chosen so as to reproduce the number density of rich clusters as observed today, which is given by the rms mass variance $`\sigma _8=0.6`$ (Eke, Cole & Frenk (1996)). The initial redshift was $`z=67`$ and the box size 50 Mpc (all length scales are quoted in real units). To ensure the collisionless nature of the dark matter does not become contaminated by two-body forces, the two-body relaxation should be longer than the Hubble time. Thomas and Couchman (1992) show that for a uniform distribution of particles (of mass m, and softening $`ϵ`$) within a spherical volume of radius $`R`$ and with a velocity dispersion $`v^2\gamma GmN/R`$ ($`\gamma `$ is a constant dependent upon the characteristics of the velocity distribution) the two-body relaxation time is $$t_r\frac{\gamma ^{3/2}N^{1/2}}{6\mathrm{ln}(R/ϵ)}\left(\frac{R}{ϵ}\right)^{3/2}t_2,$$ (10) where $`t_2=(ϵ^3/Gm)^{1/2}`$ is the minimum time-scale for interactions of particles under gravity, with an effective impact parameter $`ϵ`$. Utilising the velocity dispersion for an isothermal sphere and modifying the $`t_2`$ parameterization of TC92, equation 10 may be rearranged to give $$\frac{t_r}{t_0}0.02\frac{(R/ϵ)^{3/2}}{ln(R/ϵ)}N^{1/2}\left(\frac{ϵ}{4\mathrm{kpc}}\frac{50\mathrm{Mpc}}{box}\right)^{3/2}\left(\frac{N_p^{eff}}{256^3\mathrm{\Omega }_p}\right)^{1/2}.$$ (11) Ideally the ratio $`t_r/t_0`$ should be greater than one, although since the simulation evolves to $`z=1`$, values of $`0.5`$ should be acceptable. The formula is minimized when $`R=\mathrm{e}^{1/2}ϵ`$ (i.e. close to the softening), and at this radius it is found that $`t_r/t_00.08N^{1/2}`$. This implies that within the softening volume, $`N>10`$ is neccesary to avoid two-body relaxation, although if the simulations were integrated to $`z=0`$, the criterion would be $`N>30`$. The effective resolution of the high resolution region (which is 6.25 Mpc in diameter) is $`256^3`$, which yields a mass resolution of $`4.6\times 10^8`$ $`\mathrm{M}_{}`$ in the dark matter, $`5.1\times 10^7`$ $`\mathrm{M}_{}`$ in the gas (reducing to $`2.6\times 10^7`$ $`\mathrm{M}_{}`$ after the creation of the first star particle) and $`2.6\times 10^7`$ $`\mathrm{M}_{}`$ in the star particles. Clearly the mass resolution remains low, with a $`10^{11}`$ $`\mathrm{M}_{}`$ galaxy (in baryons) being represented by approximately 4,000 gas and star particles, assuming an equal division of both. Nonetheless, this resolution is sufficient to give a reasonable indication of the performance of different algorithms in a cosmological environment. The total baryonic mass in the high resolution region is approximately $`8\times 10^{11}`$ $`\mathrm{M}_{}`$. This is a consequence of attempting to keep the boundary of the second mass hierarchy a sufficiently long way from the object of interest, and choosing a sufficiently large box size to provide a reasonable representation of tidal forces. Since the simulated disk will was evolved over 3 Gyr, and the simulation had comparatively low resolution, shear-correction was applied to the artificial viscosity. The first simulation conducted was a low resolution $`128^3`$ dark matter simulation using the parameters given. From this simulation, candidate halos for re-simulation were extracted at a redshift of $`z=1`$. Due to wall-clock limits on simulation time, it was decided that $`z=1`$ was the most appropriate time to stop the simulation. Whilst the dark matter run could have be continued to $`z=0`$, this is prohibitively expensive for the high resolution hydrodynamic runs, since it requires close to 15,000 time-steps. The chosen halo had a mass of $`2.7\times 10^{12}`$ $`\mathrm{M}_{}`$ and is thus comparatively large. Re-simulation showed that it corresponds to the halo of a merger event of two galaxies with a combined bayonic mass of $`1.8\times 10^{11}`$ $`\mathrm{M}_{}`$. ### 5.3. System evolution without feedback Without feedback, the system follows the ubiquitous cooling catastrophe picture. Baryons condense in the halos and rapidly radiatively cool due to their high density. A disk galaxy is formed in the center of the high resolution region, with a (baryonic) mass of $`1.2\times 10^{10}`$ $`\mathrm{M}_{}`$. The disk exhibits a (visibly striking) cutoff in particle density at a radius of 8 kpc, and similarly, the vertical distribution of the disk falls off abruptly above and below 1.5 kpc of the equator. A double exponential fit of the gas density profile (see section 5.6) clearly displays the rapid fall-off in density with radius beyond 8 kpc. Star formation proceeds rapidly due to the high density, and is initially concentrated in the nucleus of the disk (which contains 40% of the baryonic mass and has a 0.6 kpc diametermuch smaller than the softening length). Stars formed in the nucleus diffuse away from it, forming a stellar bulge approximately 3.0 kpc in diameter (compare the radial density profiles in figure 15). Due to the low resolution, the hierarchical formation of the disk is represented poorly, with only a handful of progenitors merging to form the disk. At late times $`z=1.09`$, the disk exhibits a number of features that have been observed previously. There is a deficit in angular momentum, with the specific angular momentum of the baryons corresponding to that of an elliptical system for the given mass scale. Consequently the disk has a small radial extent. At $`z=1.01`$ the disk undergoes a major merger with another system of mass $`6\times 10^{10}`$ $`\mathrm{M}_{}`$ (in baryons) at a speed of 300 km s<sup>-1</sup> relative to the center of mass frame for the major disk. As is generally observed in simulations with stellar and gaseous components, the resulting morphologies of the gas and stars differ significantly. The gas cores merge, creating a very dense core while the stellar components merge, producing ‘shells’ as observed in ellipitical galaxies (Quinn 1984). A tidal tail is also produced during the merger and is populated by both gas and stars. ### 5.4. Star Formation Rate Unfortunately, even though the SFR normalization was adjusted to 0.025, the SFR in the simulations appears to be somewhat low. Although the plots in figure 13 show SFRs in excess of 70 $`\mathrm{M}_{}`$yr<sup>-1</sup>, it should be noted that this value is integrated over $`8\times 10^{11}`$ $`\mathrm{M}_{}`$. Diagnostics from the simulation indicate that of this mass, $`6.57\times 10^{11}`$ $`\mathrm{M}_{}`$ is not involved in star formation ($`T>30000`$ K or $`\rho <\rho _{sf}`$). Beyond the main disk and the merger companion (a combined baryonic mass of $`1.8\times 10^{11}`$ $`\mathrm{M}_{}`$ of which 60% is in star forming regions), tertiary halos contribute only $`2.4\times 10^{10}`$ $`\mathrm{M}_{}`$ of star forming matter. Hence, the bulk of the SFR is derived from the main disk and its major companion. It should be emphasized that the halo correspondence between simulations is not perfect, but given the difficulty in accurately calculating the cooling rates at low resolution, and the highly non-linear nature of the dynamics, this is not surprising. Well-defined halos, i.e. those formed with 500 or more particles, do correspond well, as can be seen in the radial plot in figure 19. There are also small synchronization errors ($`10^5`$ years) between the analysed time-step outputs. To examine the effect of changing parameters a number of auxiliary simulations were run, the details of which are discussed in section 5.8. ### 5.5. Effect of feedback on SFR and morphology The most noticeable difference in the ensemble is that the temperature smoothing version does not lead to a significantly different final structure. This is contradictory to the isolated results where temperature smoothing is seen to promote violent winds and disk disruption. The reason for this is that the density of the first objects is so high, Fig. 14.—Radial temperature profile for the NF, SPa and ESa runs. The error bars denote 1$`\sigma `$ variation about the bin mean, with the data plotted in Lagrangian bins of size 208 particles. The SPa and SPna (not shown) profiles both exhibit a higher temperature at large radii (see text). Of the remaining algorithms, all follow profiles similar to the ESa and NF simulations. $`n_H>1\mathrm{cm}^3`$, that the cooling time ($`0.1`$ Myr) is short enough to remove the SN energy within a time-step, unlike the isolated simulation where the resolution is high enough to allow a reduction in density due to expansion of the feedback region and consequent reduction in the cooling time. To test what happens when the feedback energy is allowed to persist, the TS simulations were run with the adjusted cooling mechanisms. As expected, these simulations produced more diffuse structures. At z=1.09, the morphology of the major disk was examined. Without exception, all the simulations produced a disk with a clearly defined cutoff radius of $`9.2_{1.6}^{+2.1}`$ kpc ($`>2h_{min}`$). This result is the same as the NF run. However, models including feedback were fatter at the disk edge, (the TSa run with a thickness of 5 kpc, being 30% wider than the NSF run). ‘Bubbles’ were noticeable in the disks, more so in the ESa run than others because the TSa and TSna runs were already quite diffuse. Feedback did not change the radial extent of the disk which suggests that it must be determined largely by the dark matter potential. It cannot be due to the central concentration of baryons, since the TSa and TSna runs effectively destroy this concentration yet still have approximately the same radius. While the internal structure of the major disk was not significantly different across all simulations, that of the merged system was. For the NF simulation, the gaseous cores were much more tightly bound than those in the TSa and TSna runs, but were largely similar to those in the ESa, Esna, SPa and SPna runs. In particular, because the gas cores are sufficiently inflated in the TSa and TSna runs, the gas undergoes a smooth merger, and for the TSa run the feedback is sufficient to produce a disk as the result of the merger. Note that the stellar components evolve in a similar fashion though, producing shells, and a widely dispersed final stellar structure. The SFRs for the main simulations are plotted in figure 13. The upper left panel shows the results for the simulation without feedback, and gives an illustration of the smoothing effect of the 160 step running average used to smooth the data. All algorithms agree on the early SFR, which reaches 1 $`\mathrm{M}_{}\mathrm{yr}^1`$ at $`z=3.9`$, since sufficient time must pass for the star mass of particles in the highest density regions to reach the mass threshold for creating a star particle (the first star particles are created at $`z3`$). At late times, the merger causes a strong star burst which is visible in all of the SFRs, albeit somewhat suppressed in the simulations with strong feedback. The relative effect of the different feedback algorithms was compared by calculating the reduction in the cumulative mass of stars at $`z=1.09`$, as a percentage relative to the no feedback run (total $`9.7\times 10^{10}`$ $`\mathrm{M}_{}`$). The algorithms with the most significant effects are, in order, TSa (25%),TSna (23%), SPa (18%) SPna (13%) and TS (10%) while the energy smoothing variants ESna (3%), ESa (2%) have comparatively little effect on the SFR. At earlier, epochs, in particular shortly after $`z=2`$, both the TS and SP runs have a significantly higher reduction in the SFR. For example at $`z=1.6`$ the SPna run has an SFR only 20% of the NF run. As in the isolated simulations, the SP algorithms eject particles due to asymmetry in the local distribution of particles and the subsequent reduction in the gas density is the main source of the SFR reduction over the energy smoothing variants. ### 5.6. Halo profiles In view of the results from the isolated simulations, the halo structure was examined to see if there was any difference between algorithms. Figure 14 compares the gas halo temperature for the NF, ESa and SPa runs. The higher temperature seen at the edge of the SP profile (beyond 200 pc), is difficult to attribute just to ‘hot’ particles being ejected to that radii. Tracing the orbits of a number of ejected particles showed that the largest distance they achieve from the core is 150 kpc. It is noticeable that at a radius of 150 kpc, the temperature of the SP feedback halos is higher than that of the others. A plot of the radial pressure showed that beyond 200 kpc, the pressure in the SP halos is a factor of two higher than in the other runs. A plot of the cumulative density versus radius confirms that more of the gas lies at large radii (beyond 200 kpc) for the SP feedback. This indicates that the particles ejected from the disc by SP feedback are acting like a piston on the outer regions of the gas halo, subsequently leading to higher temperatures in the infalling matter at large radii (since the gravitational compression remains dominated by the dark matter). The density profiles for the baryons can be fit by a double exponential, with the break between the two profiles occurring at the edge of the disk. An argument can be made that the presence of the gas/stellar core suggests that the disk should also exhibit a double profile; however the structure is sub-resolution. In particular, when the smoothed density is examined (which is used in the SFR calculation), there is very little difference between simulations. A summary of least squares fits for the inner and outer parts of the density profiles is given in table 3. The fitting was somewhat arbitrary since the break between the fits is decided by eye. Note that for the TS Fig. 15.—Density profile for the NF run. Spherical Lagrangian bins (52 particles) have been used to bin the star and gas data. Clearly the gas and stellar data are comparatively similar. The stellar bulge has roughly constant density and extends to a radius of 1.5 kpc, the gas nucleus extends only just beyond 0.5 kpc. The SPH density data is a radial binning of the raw SPH density values which, because of smoothing, do not increase to the exceptionally high values seen in the radially binned stars and gas. Further, it reflects the 2-dimensional density better than the spherical bins. The self-gravity line corresponds to 0.4 times the spherically binned dark matter density. variants, this was particularly difficult since the transition from disk to halo is less clear, i.e. the density curve is smoothly decreasing as opposed to a sharp discontinuity visible in the other data sets. The inner fits, which give an effective scale length for 3-dimensional baryon distribution in and about the disk, are broadly similar and are given by $`s_L=0.75_{0.16}^{+0.25}`$ kpc (ignoring the auxiliary simulations presented in section5.8). Note that the gas cores tend to distort the fits toward shorter scale lengths and the three-dimensional profile underestimates the scale length that would be interpreted from a surface density plot. There is a clear trend for the adiabatic feedback schemes to have longer scale lengths than the non-adiabatic versions. This is to be expected since the adiabatic feedback keeps the gas more diffuse. The outer fits are more problematical since satellites severly distort the radially averaged density profile. Given the sensitivity of the slope to these perturbations, it is difficult to draw detailed conclusions from these data. The gas and stellar density profiles are broadly similar since the stellar disk evolves out of the gas. Star particles that are formed within the dense central gas core eventually orbit at a larger radii than the parent particle since they are not affected by the viscous forces felt by the gas. A comparison of the gas to stellar density profile is shown in figure 15. The smoothed gas density (i.e. the SPH density) is remarkably similar across all simulations indicating that the SFR should be similar (modulo the effect of feedback events). The clear rise in the density at small radii is a signal of the gas core, albeit at sub-resolution scales. For the TSa/na runs this core was removed due to Fig. 16.—Radial velocity dispersions in the gas disk. The minimum smoothing length $`h_{min}`$ and disk edges are marked for clarity. In contrast to the isolated simulations, there appears to be no correlation between more violent feedback producing higher velocity dispersion. This is partially due to the fact that merging dwarfs produce a far higher velocity dispersion than feedback, and also the disks are not well resolved. the strong feedback. For the SP runs, the ejection of particles also lowered the core mass. The ES runs were incapable of inflating the core once formed. The density profiles for the dark matter differ little from simulation to simulation ($`r_{200}`$ differs across all simulations by only 1%). At least with this mass resolution, there is no evidence for feedback being capable of rearranging the dark matter structure (Navarro et al. (1996)). ### 5.7. Rotation curves and angular momentum Figure 17 displays the (Plummer softened) rotation curves, $$v_c^2(r)=\frac{GMr^2}{(r^2+ϵ^2)^{3/2}},$$ (12) and particle tangential velocities for the run without star formation compared to the ESa, SPa and TSa runs. For all simulations, the tangential velocities rise more sharply than the rotation curve. However, the initial slope of the rotation curve is dominated by the softening parameter and a 12% reduction in the softening length is enough to fit the tangential data, hence this should not be considered a significant discrepancy. The 300 km s<sup>-1</sup> peak of the rotation curve is consistent with the mass of the disk and halo, although the SPa run is slightly lower since gas has been ejected out of the disk into the halo. The TSa run shows significantly increased dispersion in the tangential velocities because of the feedback, but does not have a larger disk diameter (in keeping with the similar scale lengths found in the analysis of the density profiles). Since particles involved in feedback regions tend to be ejected vertically from the disk, i.e. preferentially into low density regions, a large tangential velocity dispersion can arise. This is seen most clearly in the TSa plot. In view of the rotation curves being similar across all simulations, it would be expected that the velocity dispersions should also exhibit similar profiles. A comparison plot of the NSF run compared to the SPa, TSa and TSa-SG-2c\* run (note this run has double the fiducial SFR and no self-gravity criterion, see section 5.8) is shown in figure 16. As expected, the profiles are broadly similar with a maximum difference between the plotted runs of 20 km s<sup>-1</sup> at the disk edge. It is interesting to note that the TSa run has a large tangential velocity dispersion, as is visible in figure 17, but a comparatively low dispersion in the radial direction. This is a result of feedback preferentially ejecting particles vertically, in turn boosting the tangential velocity component more than the radial. The data for the NF run (not shown) are dominated by an ongoing merger which introduces a very large dispersion (120 km s<sup>-1</sup>) at the edge of the disk, far larger than that produced by any feedback. The NF and ESa runs also show the effect of the merger, with peaks in the data at 4.2 kpc and 3.6 kpc respectively, corresponding to the position of the strongest perturbation within the disk. The SPa and TSa runs are less affected by the merger since the dwarf has been reduced in mass by the stronger feedback. It is clear that the low resolution in the disks and the complications of ongoing mergers make it difficult to draw conclusions from this data. To see how much angular momentum has been lost by the disk, the specific angular momentum of the cores is compared to that of the dark matter halo (within $`r_{200}`$) in figure 18. The angular momentum for the halo gas (the gas for which $`\delta <2000`$ within $`r_{200}`$) and that of the stellar component of the disk are also shown. For all simulations, the disk system shows a deficit of specific angular momentum when compared with the dark matter. By breaking the disk into its stellar component and gas component, it becomes clear that in the simulations with feedback that are shown (TSa, SPa and TSa-SG-2c) there is a trend toward higher angular momentum values for the gas disk. The highest value, that from the TSa-SG-2c simulation, just falls within the disk region of the parameter space. However, the stellar disks all fall in the elliptical region of the parameter space, and the effect of feedback on them is small. Note that the purely gaseous run also sits in the elliptical region and the gas disk in the no feedback run has marginally higher specific angular momentum than the stellar component. The NF run is misleading, since a merger is going on at the edge of the disk leading to higher angular momentum values as compared to the other feedback runs (although by the end of the merger the opposite Fig. 18.—Specific angular momenta versus mass for different components of the system for a number of different feedback algorithms. The filled polygons plot the angular momentum of the dark matter within $`r_{200}`$, the open polygons that of halo gas (all gas that does not fall above $`\delta =2000`$), pointed stars that of the gas in the main disk ($`\delta >2000`$) and finally the centrally connected stars show that angular momentum of the stellar component of the disk. The runs with feedback show a small but noticeable trend toward higher angular momentum values for the gas disk component (contrast with the NF run). Both dark matter and gas halo values are in broad agreement as expected. will be true due to core-halo interaction). For all simulations, the angular momentum of the halo gas is larger than that of the dark matter since the dark matter plot includes the contribution of the dark matter core, which has little angular momentum but a significant amount of mass. It is clear from the plot that if the halo gas were to fall smoothly onto the disk, then it should be possible to form a disk with an angular momentum value midway between that of the halo gas and disk system. Note infalling satellites may still disrupt the disk and cause still further angular momentum loss. The z-component of the specific angular momentum L is shown in figure 19. If significant angular momentum loss occurs as a result of disk formation then a larger proportion of the disk mass should lie beneath the line formed by $`|𝐫\times 𝐯_{\mathrm{𝐜𝐢𝐫𝐜}}|`$. Since this is not the case, it is clear that at $`z=1.09`$ there has been little angular momentum loss (within the disk) due to bar formation. However, all the disks are deficient in angular momentum relative to the halo, since they have been formed in a hierarchical process which is subject to core-halo angular momentum transport. Remarkably, the $`X_2(R)`$ stability parameter (Toomre (1981); Binney & Tremaine (1987)) plots are all similar, with all the disks achieving the $`X_2(R)3`$ stability requirement just beyond the 4 kpc softening radius. As was shown by Dominguez-Teneiro et al. (1998), it is more than likely that if the baryonic mass is redistributed into an exponential disk, then the $`X_2(R)`$ parameter for this system does not achieve stability until a much larger radius, i.e. the cores provide disk stability. Note that the kink in the $`X_2(R)`$ plot for the NF run is due to the merger previously discussed and it does not affect the disk stability greatly (the kink drops to 2 at 6.5 kpc but quickly returns to stable values). ### 5.8. Auxiliary simulations To adequately examine the parameter space of these simulations and also determine the effect of SPH algorithm changes would take an excessively long time. Alternatively, by conducting a few auxiliary simulations, much can be learnt about the outer edges of the parameter space. To understand what happens when the SPH algorithm is changed, in particular in relation to the treatment of high density regions, it is simple to contrast to one of the previous simulations run with the same parameters. As indicated in previous sections, to determine the effect of removing the self-gravity criterion, the ESa simulation was rerun without it. This simulation is denoted ESa-SG. Also the effect of doubling the star formation rate normalization was tested in a simulation denoted ESa-2c\*. Since most of the simulations conducted appeared to be relatively unchanged by the introduction of feedback, a simulation with extremely violent star formation and feedback was run (twice the fudicial SFR normalization with TSa feedback and also without the self-gravity criterion). This simulation, denoted TSa-SG-2c, is not particularly realistic (star formation begins very early and the feedback is over-efficient) but it does provide an excellent guide to the limits of what feedback can accomplish. Because of the strong feedback, the formation of dense gas cores was all but prevented, with the exception of very large $`>5\times 10^{10}`$ $`\mathrm{M}_{}`$ systems. Hence it was possible to integrate the system to later times since the SPH algorithm did not suffer significant slow down. An ESa-SG-2c simulation was also run to contrast with the temperature smoothing version. As an alternative to smoothing over all the particles within the minimum smoothing length, $`h_{min}`$, literature on galaxy formation (this appears to be implied in Navarro & White (1993); Bate & Burkert (1997)) suggests that one may continue search over $`N_{smooth}`$ neighbour particles. Such a procedure places a direct limit on the maximum resolved density: the volume normalization is set by $`h_{min}`$ and the summation over neigbours is limited to $`N_{smooth}`$ particles. In turn, this sets a bound on the maximum SFR per particle and consequently within the system as a whole. It also sets a bound on the cooling rate. The reasons for making this change are primarily related to efficiency: if a simulation smooths over all the particles within $`h_{min}`$ it can exhibit a severe slowdown. For example, nearly all the simulations without this adjustment slow down by a factor of 7 from start to finish. Hence to examine the effect of making this change, the ESa simulation was re-run with this high density treatment. This simulation is denoted ESa-nav. It is interesting to note that this method can be made to almost exactly reproduce the density values calculated in a simulation with $`h_{min}0`$, a volume normalization is the only thing that need be applied. In the ESa-SG run, the effect of removing the self-gravity criterion is that star formation begins at a very early epoch (1 $`\mathrm{M}_{}\mathrm{yr}^1`$ at $`z=8.3`$, see figure 20). It is initially confined to a few particles (recall the $`.𝐯`$ criterion must also be fulfilled) leading to an SFR of 0.2 $`\mathrm{M}_{}\mathrm{yr}^1`$. During later evolution, the SFR is only marginally lower than that of the ESa run, the largest difference being 10 $`\mathrm{M}_{}\mathrm{yr}^1`$ at $`z=2.1`$ (a 12% difference). Further, the disk and halo structure are comparatively similar, as measured by density and temperature profiles, and the dense gas core is still formed. The plot of the SFR in the ESa-2c simulation (figure 20) shows that doubling the SFR normalization leads to a stronger initial star burst as the gas overcomes the self-gravity criterion, but does not produce an SFR that is exactly double that of the standard simulations. The peak SFR of 120 $`\mathrm{M}_{}\mathrm{yr}^1`$ at $`z=2.1`$ is 56% higher than that in the ESa run (77 $`\mathrm{M}_{}\mathrm{yr}^1`$). The SFR is not simply doubled because of both increased feedback and the finite amount of gas available for star formation. Notably the dense gas core was still formed, showing that even with the increased SFR, ESa is still incapable of producing a significant effect on morphology. As expected, the TSa-SG-2c simulation leads to markedly different results than any of the previous simulations. Because of the extreme feedback and consequent absence of small progenitors, the disk assembly process is very smooth. There is essentially no formation of a gas and stellar nucleus within the disk. Star formation begins at the same epoch as the ESa-SG run, albeit at a higher rate due to the increased SFR normalization which yields 1 $`\mathrm{M}_{}\mathrm{yr}^1`$ at $`z=9.1`$. At $`z=6.2`$ the SFR reached 15 $`\mathrm{M}_{}\mathrm{yr}^1`$ and a peak of 52 $`\mathrm{M}_{}\mathrm{yr}^1`$ occurred at $`z=3.0`$. The most noticeable difference in the SFR is that at late times it falls off precipitously. At $`z=1.0`$ the SFR is 18 $`\mathrm{M}_{}\mathrm{yr}^1`$ compared to 36 $`\mathrm{M}_{}\mathrm{yr}^1`$ for the ESa run and by $`z=0.5`$ it had fallen to 5 $`\mathrm{M}_{}\mathrm{yr}^1`$. Note the decay in the SFR versus time was linear rather than exponential, This has been observed before in parameter space explorations (Thacker (1997)). The less energetic feedback provided by energy smoothing leads to the ESa-SG-2c simulation producing a much higher peak SFR, namely 80 $`\mathrm{M}_{}\mathrm{yr}^1`$ at $`z=2.7`$. In keeping with all the other energy smoothing simulations, a central gas nucleus was formed in the disk and the temperature and density profiles remain similar to the ESa run. At $`z=0.5`$, the TSa-SG-2c disk was analysed to see if the accretion of the halo gas had indeed allowed the formation of a disk without an angular momentum deficit. By this epoch, the disk had grown to a diameter of 12.0 kpc, which is 17% larger than the value from $`z=1.09`$ and the mass had increased by 30% to $`1.39\times 10^{11}`$ $`\mathrm{M}_{}`$. $`r_{200}`$ had grown to 285 kpc. The ratio of the specific angular momenta for the gas core and dark matter, increased to 0.40 (20% increase). Visualization of the system shows that the ratio cannot increase significantly as there are very few cold gas clumps within the dark matter halo available for accretion. Most have been blown apart by feedback. Note that at the center of the halo the gas density is approximately $`n_H10^3`$ cm$`{}_{}{}^{}3`$ and the temperature is over $`10^6`$ K, leading to a cooling time of greater than $`4\times 10^9`$ years. Hence very little of this gas, which has a large specific angular momentum, can cool on to the gas disk before $`z=0`$. The results from the ESa-nav simulation are very different. The peak SFR is 38 $`\mathrm{M}_{}\mathrm{yr}^1`$ and it occurs at a much later time than the rest of the simulations ($`z=1.1`$). Also star formation begins slightly later than that seen in the simulations without the self-gravity criterion (1 $`\mathrm{M}_{}\mathrm{yr}^1`$ by $`z=6.9`$) which is significantly earlier than the standard simulations (1 $`\mathrm{M}_{}\mathrm{yr}^1`$ by $`z=3.9`$). The reduced SFR is due to the density values calculated by this method being lower and not due to any increased effect of feedback. The change in the epoch at which the self-gravity criterion is overcome is due to the change in the search radius. In the center of the halo the dark matter profiles have a much shallower density profile than the baryon cores. When the neighbour search is conducted over the reduced radius only the core is sampled rather than the full 4 kpc radius, i.e. the sharp decline in baryon density at the disk edge is ignored. The disk formed is broadly similar to that in the ESa run, although it has a slightly larger baryon core, a thinner structure (2 kpc thick) and a smaller radius. As was expected, the wall-clock per time-step slowdown was less severe for this code and it was roughly double the speed of the standard implementation. Note that this is an underestimate of the efficiency since the $`h`$-update algorithm did not accurately calculate the required change in the search radius for regions where it was less than $`2h_{min}`$, see Thacker et al (1998) for a full discussion. The algorithm preferentially smooths over too many particles, and an accurate calculation would require a very short time-step. ## 6. Summary and Discussion This paper details a study of a number of different feedback algorithms, comparing the effect on high resolution isolated systems and low resolution hierarchical simulations. The parameter space of the model was explored as were the effects of small changes in the hydrodynamic solver. Principal conclusions follow: 1. As would be expected on energy budget grounds, the temperature smoothing (TS) feedback algorithm has the most impact on structure formation. Single particle (SP) feedback has a less significant but still noticeable effect and it produces a distinct change in the temperature profile of the halo at large radii. Energy smoothing (ES) is the least effective of the three fundamental mechanisms. 2. Feedback can be shown to have a large effect in systems that are well resolved in terms of particle number. In particular, the Milky Way and NGC 6503 prototypes are strongly affected by feedback even though the softening lengths were chosen to be comparable with those of cosmological simulations. Both ‘blow-out’ and ‘blow-away’ can be produced, depending upon the feedback algorithm. 3. The NGC 6503 prototype is more strongly affected by feedback than the Milky Way prototype, i.e. feedback has more effect on low mass systems, as expected. In the NGC 6503 model it is possible to differentiate between the cooling mechanisms. 4. In hierarchical simulations even an excessive amount of feedback, TSa-SG-2c for example, produces little effect on the properties of large ($`>10^{11}`$ $`\mathrm{M}_{}`$) disk galaxies at early epochs. Rotation curves and density profiles remain broadly similar. 5. Although small at the mass scale probed by the hierarchical simulations, there is a distinct trend toward higher specific angular momentum values in the disk with increased feedback. At lower mass scales, equivalently at higher resolution, the effect should be more noticeable. 6. The revised cooling mechanism has little effect for the energy smoothing algorithm since insufficient energy is input, and consequently, the estimated density does not lower the cooling rate significantly. Conversely, for single particle feedback and temperature smoothing, the energy input is more than sufficient to force the cooling rate from the estimated density to be much longer than the local one. 7. Morphologies remain broadly similar in hierachical simulations although feedback can reduce the SFR, compared to runs without it, by over 25% at z=1. At earlier epochs reductions of significantly over 50% were seen. Note that the reduced SFRs continue to offer similar peaks and troughs albeit at a lower overall normalization. 8. The introduction of a self-gravity criterion to prevent catastrophic star formation at high redshifts does not lead to significantly different structure formation provided that the SFR normalisation is set within reasonable bounds. 9. The treatment of the high density end of the SPH solver can produce an enormous difference in the SFR. By reducing the neighbour search in the high-density regions the SPH density becomes lower and consequently so does the SFR. Also the self-gravity criterion is overcome earlier since as the search radius is reduced the baryonic cores have a higher weighting. At the resolution provided by the cosmological simulations hierarchical merging is not adequately modeled. Only a handful of progenitor halos merge to form the major disk system. That all of the simulations, including those with the exceptional feedback provided by temperature smoothing, produce an over-compact disk is not cause for concernthe radius of the disk is determined largely by the depth of the potential well. Due to the early stage at which the simulations were stopped, it is difficult to comment on the evolution of the majority of disk systems, which are expected to accrete a large fraction of mass from $`z1`$ onwards. The one simulation that was integrated to $`z=0.5`$ gave some surprising results. If feedback is sufficiently strong to reduce the angular momentum problem for the largest systems then there are insufficient halos at later times to accrete on to the disk. Although it is a significant jump to go from this result to the idea that feedback was stronger at higher redshifts it certainly seems appealing. Silk (1998) discusses variations in the IMF over time. One particularly important area that has not been examined is the effect of resolution. For star formation algorithms based upon density values, there are many issues to be considered. In particular, in a hierarchical cosmology, star formation will be higher and begin at earlier epochs with increased resolution. The effect on feedback is less clear but by increasing the mass resolution in simulations the escape velocity of the first halos is reduced. Since the gas temperature following a feedback event is not reduced (with increased resolution), heated gas will tend to orbit higher in the potential well and be more diffuse. It was noted that in the simulations with energy smoothing (even those with a high SFR) the formation of a dense central core was unavoidable. This might be partially associated with the delay between the first star formation and the first feedback. During this time it may be the case that gas can accumulate above the mass threshold at which it can be blown out. A probablistic star formation algorithm might improve this matter somewhat, although in the limit $`N\mathrm{}`$ these algorithms should converge. However, it should be acknowledged that the dark matter halos in the sCDM picture have a strong central concentration and that the core accumulation may be a result of this. As is widely known, core-halo angular momentum transport presents great problems for the sCDM picture. An important question is what is the extent of the problem for merging halos of unequal mass? It is unclear whether the limiting case of accreting a large number of low mass halos will lead to a system that has not lost a significant proportion of its angular momentum. Since the internal angular momentum of the final object is carried by the orbital angular momentum of the progenitors, there is reason to believe that it should be possible. However, higher resolution SPH studies (Thacker and Couchman, forthcoming) indicate that it is almost impossible to avoid the formation of large objects from mergers of smaller ones without the angular momentum problem coming in to play – there are too many ‘medium sized’ halos that collapse within the galaxy halos. It is also possible to make insights without relying upon high resolution since the power spectrum relevant to galaxy formation is the approximately scale invariant $`P(k)k^2`$ (cooling times are shorter for the smaller halos, and the power spectrum is tilted more toward $`P(k)k^3`$). The simulations thus performed give a good indication of the properties of the low mass progenitors of galaxies – they too should suffer from an angular momentum deficit. However, since this is a problem for the internal angular momentum and not the orbital, the effect on the final object may not be a significant problem. In light of this argument, it is seen that smooth infall in simulations occurs as a result of a lack of resolution. Indeed the idea of smooth infall in sCDM is largely a misnomer (at least without some kind of feedback mechanism). In simulations, the minimum mass scale effectively keeps the gas supported against the collapse that would normally ensue given higher resolution. Thus the disks that are formed in SPH simulations are a result of this lack of numerical resolution, and ideally detailed convergence studies should be undertaken. A number of authors have already hinted at this problem (Evrard et al. 1994; Weinberg, Hernquist & Katz 1997, for example), but as yet no systematic attempt has been made to deal with it or its effects. This problem is revisted in a paper in preparation. The large variations in SFRs that may be produced by small changes in the parameter space remain a significant concern. The treatment of the high density end of the SPH solver is of particular importance, since it can strongly affect the SFR calculated, and further the treatment is not performed in the same manner by all research groups. The development of a standard test case for dynamical star formation algorithms is desireable, the facilitating the comparison of different research results. Unfortunately, it is far from clear what kind of a test case should be adopted. Simple rotating cloud collapse models are not adequate since they provide no representation of the hierarchical formation process or the effect of tidal fields. Also, since some algorithms are grid-based and some are particle-based, it may be necessary to adopt a suite of similar test cases. McGaugh (1998) presents a survey of a number of reasons why sCDM is unable to form low surface brightness (LSB) galaxies similar to those observed. Nonetheless, it is still instructive to compare the SFRs calculated with those from deep observations of star forming galaxies. Studies of the global SFR, when adjusted for dust extinction, suggest that there is no decline between $`z=1`$ to 4.5. Consequently, the self-gravity criterion imposed, causing an abrupt turn on of star formation at $`z=4`$, does not fit the data, especially the observations of Chen et al. (1999) suggesting star formation at $`z7`$. This should not be over-interpreted since it is a result of a lack of resolution, adding higher resolution would move the onset of star formation to progressively earlier times. Turning off the criterion allows star formation to proceed very early on, at $`z=8.3`$, but it also allows star formation to occur in regions where the dark matter may still be dynamically dominant, which is undesirable. Using the continuum UV flux, the DEEP survey of the Hubble Deep Field derives SFRs from 0.14 $`\mathrm{M}_{}\mathrm{yr}^1`$ to 24.92 $`\mathrm{M}_{}\mathrm{yr}^1`$ for $`q_0=0.05`$ and $`q_0=0.5`$ values are 11 times lower (Lowenthal et al (1997)). Note that these values are not corrected for dust extinction. At $`z=3`$ the SFRs calculated (30-70 $`\mathrm{M}_{}\mathrm{yr}^1`$) are higher than those found, although the selected sample are categorized as “large dwarf spheroidals” and, in contrast, at $`z=3`$ the simulated galaxy already has a mass of $`7.2\times 10^{10}`$ $`\mathrm{M}_{}`$. More recently estimates of SFRs derived from H$`\beta `$ emission, for a sample of $`z=3`$ galaxies, have shown widely diverging values compared to the estimate from the UV flux (with both fluxes being uncorrected for dust extinction). The H$`\beta `$ values are larger, by as much as a factor of 7, yielding SFRs in the range 20-270 $`\mathrm{M}_{}\mathrm{yr}^1`$ (Pettini et al. (1998)). It is difficult to see how the models presented can be tuned to produce SFRs in the region of 270 $`\mathrm{M}_{}\mathrm{yr}^1`$ since it would require an SFR normalization far beyond what we believe is the realistic parameter space. A forthcoming paper examines the effect of higher spatial resolution on the derived SFR. The authors thank NATO for providing collaborative research grant CRG 970081 which helped facilitate this research. Useful discussions with Drs Fabio Governato and Frazer Pearce are acknowledged. RJT was supported by a Dissertation Fellowship from the University of Alberta while this research was conducted. HMPC ackowledges support from NSERC Canada.
warning/0001/quant-ph0001044.html
ar5iv
text
# Untitled Document Quantum key distribution based on Greenberger-Horne-Zeilinger state Guihua Zeng National Key Lab. on ISDN, XiDian University, Xi’an 710071, P.R.China Abstract An unsymmetrical quantum key distribution scheme is proposed, its security is guaranteed by the correlation of the Greenberger-Horne-Zeilinger triplet state. In the proposed protocol, the distribution of quantum states are unsymmetrical. This unsymmetrical characteristic makes the transmission qubits (except the loss qubits) be useful. Sequentially the proposed protocol has excellent efficiency and security which are very useful in the practical application. PACS: 03.67.Dd, 03.65.Bz, 03.67.-a Quantum key distribution is defined as a procedure allowing two (multi-) legitimate users of communication channel to establish two (multi-) exact copies, one copy for each user, of a random and secret sequence of bits. It employs quantum phenomena such as the Heisenberg uncertainty principle and the quantum corrections to protect distributions of cryptographic keys. Since the BB84 protocol was presented, lots of quantum key distribution protocols for two communicators have been proposed. Three main protocols are the BB84 protocol , B92 protocol and the EPR protocol . All these presented quantum key distribution protocols are provably secure against eavesdropping attacks \[5-9\], in that, as a matter of fundamental principle, the secret data can not be compromised unknowingly to the legitimate users of the channel. Currently, the quantum key distribution has entered the experimental phases. The first quantum key distribution prototype, working over a distance of 32 centimeters in 1989, was implemented by means of laser transmitting in free space . Soon, experimental demonstrations by optical fibber were set up . Recently, the transmission distance is extended to more than 30Km in the fiber and 205m in the free space . The first quantum key distribution scheme, i.e., the Bennett Brassard (BB84) scheme, was presented a decade ago. It is implemented by the four states $`\{|,|,|,|\}`$, where any of the two states $`\{|,|\}`$ and any of the two states $`\{|,|\}`$ are non-commuted. Its security relies on the uncertainty principle of quantum mechanics. The security guarantee is derived from the fact that each bit of data is encoded at random on either one of a conjugate pair of observables of quantum-mechanical object. Because such a pair of observables is subjected to the Heisenberg uncertainty principle, measuring one of the observables necessarily randomizes the other. In 1992, Bennett devised another protocol, i.e., the B92 protocol. It is based on the transmission of nonorthogonal quantum states. This protocol uses any two nonorthogonal states to implement the quantum key distribution. Its security relies on the non-distinguishability of unknown two-nonorthogonal states. Recently another quantum key distribution scheme, based on Einstein-Podolsky-Rosen (EPR) correlations, was suggested by Ekert and modified by Bennett, Brassard, and Mermin. In the Ekert’s vision, it uses the EPR pair to distribute the quantum cryptographic key, and uses the violation of the Bell inequalities to provide the secret security. In the Bennett, Bassard and Mermin’s revised vision, they use the EPR pair to distribute the quantum cryptographic key, but use the EPR correlation to provide the secret security. We describe here the modified version. In this scheme the communicator, called Alice, creats pairs of spin 1/2 particles in the singlet state, and sends the communicator, called Bob, one particle from each pair. When the two particles are measured separately the results obtained for them are correlated. Using the correlation Alice and Bob can check the eavesdropping and obtain the sharing key. In this letter, we present another method in which the communicators use the Greenberger-Horne-Zeilinger (GHZ) state to distribute the quantum cryptographic key and use the correlations of the GHZ triplet state to provide the secret security. In this scheme two particles of the GHZ triplet state are distributed to Alice and the third particle of GHZ triplet state is distributed to Bob. Obviously, the distribution of the particles and the quamtum states are unsymmetrical in the proposed protocol, which is different from the previous schemes in which Alice and Bob the same numbers particles or quantum states. This unsymmetrical characteristic brings some merits than the previous protocol, for example the efficiency approaches $`100\%`$ and the security is more higher, these characteristics are very useful in the practical application. In the proposed protocol we let Alice or the quantum source create a sequence of GHZ triplet states, and send one particle from each triplet state to Bob. When three particles are measured the results obtained for them are correlated. By using the correlation of the GHZ triplet state Alice and Bob can check the eavesdropping and obtain the sharing key. Before describing our protocol, we first investigate the correlation of the GHZ triplet state. In general, A N-particle entanglement states may be written as $$|\psi =\underset{i=1}{\overset{N}{}}|u_i\pm \underset{i=1}{\overset{N}{}}|u_i^c,$$ (1) where $`u_i`$ stand for a binary variable, $`u_i\{|z+,|z\}`$ and $`u_i^c=1u_i`$, $`|z+`$ and $`|z`$ denote the spin eigenstates, or equivalently the horizontal and vertical polarization eigenstates, or equivalently any two-level system. For $`N=2`$ they reduce to the Bell states and $`N=3`$ and $`N=4`$ they represent the GHZ states. For a general $`N`$ we shall calling them cat states. In this paper, we are interesting in the case of $`N=3`$, i.e., the GHZ triplet state. In Eq.(1) for $`N=3`$ they reduce to eight GHZ triplet states, we here use the following state $$|\psi =\frac{1}{\sqrt{2}}\left(|z+z+z++|zzz\right).$$ (2) Define the $`x`$ and $`y`$ eigenstates $$|x+=\frac{1}{\sqrt{2}}\left(|z++|z\right),$$ (3) $$|x=\frac{1}{\sqrt{2}}\left(|z+|z\right),$$ (4) $$|y+=\frac{1}{\sqrt{2}}\left(|z++i|z\right),$$ (5) $$|y=\frac{1}{\sqrt{2}}\left(|z+i|z\right),$$ (6) the GHZ triplet state can be rewritten as $$\begin{array}{cc}\hfill |\psi =& \frac{1}{2}\left[\left(\right|x+\right|x++|x|x)|x+\hfill \\ & \\ & +\left(\right|x+|x+|x|x+)|x],\hfill \end{array}$$ (7) or $$\begin{array}{cc}\hfill |\psi =& \frac{1}{2}\left[\left(\right|y+\right|y+|y|y+)|x+\hfill \\ & \\ & +\left(\right|x+|x+|x|x+)|x],\hfill \end{array}$$ (8) or $$\begin{array}{cc}\hfill |\psi =& \frac{1}{2}\left[\left(\right|y+\right|x+|y|x)|y+\hfill \\ & \\ & +\left(\right|y+|x++|y|x)|y],\hfill \end{array}$$ (9) or $$\begin{array}{cc}\hfill |\psi =& \frac{1}{2}\left[\left(\right|x+\right|y+|x|y+)|y+\hfill \\ & \\ & +\left(\right|x+|y++|x|y)|y].\hfill \end{array}$$ (10) The above decomposition demonstrates the correlation among three particles. For example, in Eq.(7) if one particle is in the state $`|x+`$ and the second particle is in the state $`|x+`$, the third particle must be in the state $`|x+`$ because of the correlation of the GHZ triplet state. By Eqs.(7-10), one may construct a lock-up table to summarize these properties of GHZ states. Table I. The correlation results of the GHZ triplet states | Particle 1 | $`|x+`$ | $`|x`$ | $`|y+`$ | $`|y`$ | | --- | --- | --- | --- | --- | | Particle 2 | $`|x+`$ | $`|x+`$ | $`|x+`$ | $`|x+`$ | | Particle 3 | $`|x+`$ | $`|x`$ | $`|y`$ | $`|y+`$ | | Particle 2 | $`|x`$ | $`|x`$ | $`|x`$ | $`|x`$ | | Particle 3 | $`|x`$ | $`|x+`$ | $`|y+`$ | $`|y`$ | | Particle 2 | $`|y+`$ | $`|y+`$ | $`|y+`$ | $`|y+`$ | | Particle 3 | $`|y`$ | $`|y+`$ | $`|x`$ | $`|x`$ | | Particle 2 | $`|y`$ | $`|y`$ | $`|y`$ | $`|y`$ | | Particle 3 | $`|y+`$ | $`|y`$ | $`|x+`$ | $`|x`$ | In any column of the table I we assume the result of particle 1 to be determined, and consider the possible results of particles 2 and 3. The table I shows two properties of the GHZ triplet state: i) By the result of any of the three particles, one can determine whether the other two results of the particles are the same or not and also that he (she) will gain no knowledge of what their results actually are, if he (she) knows what measurements have been made for the other two particles (that is $`x`$ or $`y`$). ii) From table I it is clear that allows two parties jointly, but only jointly, to determine which was the measurement outcome of the third party. So if the measurement directions of the three participators are public, the combined results of any two participators can determine what the result of the third party’s measurement was. As discussed above, the GHZ state has a correlation properties that if only one particle has been measured, the states of the other two particles are still not determined. This property may be used in the quantum key distribution scheme. Let us now show how to implement the quantum key distribution scheme by the GHZ state. Suppose Alice have two particles denoted by $`P_1,P_3`$, and Bob have one particle denoted by $`P_2`$ from the GHZ triplet state. The protocol goes as follows. 1. Alice makes a random measurement on her one particle (e.g., $`P_1`$) of two GHZ particles, either in the $`x`$ or $`y`$ direction. 2. Bob makes a random measurement on his GHZ particle $`P_2`$, either in the $`x`$ or $`y`$ direction. 3. Bob sends his measurement bases $`x`$ or $`y`$ to Alice, but not the qubit values. 4. Alice measures his particle $`P_3`$ according to the measurement bases of the particles $`P_1`$ and $`P_2`$. 5. Check eavesdropping by using the correlation of the GHZ triplet state. 6. Alice judges the Bob’s result by the results of the particles $`P_1,P_3`$ according to table I, and transfers her results to be consistent with Bob’s results. 7. Alice and Bob obtain a sharing key by using the data sifting, the error correction and privacy amplification technologies. Let us now explain the above protocol in detail. In our protocol the sequence of the GHZ triplet states may be generated by one communicator, e.g. Alice or by a quantum source. For each GHZ triplet we let Alice have two particles and Bob have one particle from the GHZ triplet. The roles of the Alice’s two particles are different: one is for the key and another is for setting up the correlation of the GHZ triplet. In steps 1 Alice only measures one of two GHZ particles, the result of this particle will be used for the quantum cryptographic key. The another particle will be measured in step 4. This particle is used to judge the Bob’s result. In additional, by making use of the particle $`P_3`$ Alice and Bob do not need to discard the sent quantum bits (qubits) that have been always discarded in all the previous protocols. This is the superiority of our protocol. In step 4 Alice’s measurement on particle $`P_3`$ must base on the measurement bases of the particles $`P_1`$ and $`P_2`$. The reason is that after Alice and Bob measure respectively their particles $`P_1`$ and $`P_2`$, the state of the particle $`P_3`$ is determined by the correlation of the GHZ triplet state. If both Alice and Bob measure their particles ($`P_1,P_2`$) using the same measurement bases, i.e., $`x`$ or $`y`$ direction, Alice measures her particle $`P_3`$ along the $`x`$ direction, otherwise, Alice measures her particle using the $`y`$ measurement basis. After measure Alice gains the result of the particle $`P_3`$, which is any of the four states $`\{|x+,|x,|y+,|y\}`$. Alice and Bob check the eavesdropping by using the correlation of the GHZ triplet state. If the results of the particles $`P_1,P_2,P_3`$ satisfy any of the Eqs.(7-10), which means the results are in the table I with the correlation of the GHZ triplet state, we say the correlations among three particles are perfect. After transmission, Alice and Bob publicly compare the results of the GHZ triplet states on a sufficiently large random subset of the set of all GHZ triplets. If they find that the test subset is indeed perfectly correlated, they can refer that the remaining untested subset is also perfectly correlated, and therefore may be used for the quantum cryptographic key. The method is as follows: Bob send publicly Alice a random sub-sequence of his results, Alice compares the Bob’s results with her corresponding results. If the correlations of their results may be found in the table I, the correlation is perfect, otherwise it means the eavesdropping. For example, if the results of the particles $`P_1,P_2,P_3`$ are respectively the $`|x+,|x,|x`$, the results are perfect, if the results of the particles $`P_1,P_2,P_3`$ are respectively the $`|x,|x,|x`$, Bob’s particle must be eavesdropped by the eavesdropper or be disturbed by the noise. It needs to stress that Alice’s results must be consistent with Bob’s results for getting the raw quantum key. In step 6, Alice can judge Bob’s results by the results of the particles $`P_1`$ and $`P_3`$, but Bob gains no knowledge on Alice’s results, because Bob only has one result of the GHZ triplet state. In addition, from table I we see the most correlation results of particles $`P_1,P_2,P_3`$ are different. For example when Alice’s results of two particles $`P_1,P_3`$ are $`|x`$ and $`|x`$, according to table I the Bob’s result should be $`|x+`$, obviously, the results of the particles $`P_1`$ and $`P_2`$ are different. For obtaining a same key, Alice’s results need to be consistent with Bob’s results. We may use two methods to gain the same key. One is that Alice changes her qubits according to Bob’s results. Another is that Alice transfers her qubits to binary bits according to Bob’s results. For example, assume Bob transfers his qubits using the aforehand appointment $`|x+,|y+0,|x,|y1`$, if Bob’s result is $`|y+`$ and Alice’s result is $`|x`$, both Alice and Bob transfer their qubits to binary bits $`0`$. This way is different from the previous protocols, in which Alice and Bob have a same aforehand appointment. As the previous protocols the raw quantum key distribution is useless in practice because limited eavesdropping may be undetectable, yet it may leak some information, and errors are to be expected even in the absence of eavesdropping. For these reasons, our scheme needs to supplement some classical tools such as privacy amplification, error correction and data sifting, so we add the step 7 in our protocol. The implementation of these supplemented classic tools are the same as in the previous documents . The important point is that the efficiency of the quantum key distribution has been improved in this protocol. Because Alice and Bob randomly measure their particles $`P_1`$ and $`P_2`$ before the Alice’s measurement on the particle $`P_3`$, and Alice can measure the particle $`P_3`$ according to the measurement bases of the particles $`P_1,P_2`$, no Bob’s qubits need to be discard in the proposed scheme. This way largely improves the efficiency of the distribution for cryptographic key. Consider the quantum cryptographic system will not be perfect because of the losing qubits ($`l`$) in the measurement and in the quantum channel, in order to be left with a key of $`L`$ qubits Alice should send $`L^{}>(L+l)`$ qubits to Bob. The efficiency is $$\eta =\frac{L}{\left(L+l\right)}<100\%.$$ (11) In the previous protocol, the optimal efficiency is $`\eta ^{}=L/2(L+l)<50\%`$. Obviously, when the lossing qubits are same, $`\eta >\eta ^{}`$. From Eq.(11), we see as $`l0`$ $`\eta 100\%`$, it means that an ideal unsymmetrical GHZ protocol corresponds to the $`100\%`$ efficiency. It need to stress that here the discard particles are the same as BB84 protocol for obtaining a $`L`$ qubits because the particle $`P_3`$ is finally discard in our scheme. Let us now consider the security of this protocol. It is warranted by the correlation of the GHZ triplet state. From the table I it is clear that even if the eavesdropper knows what measurements Alice and Bob made (that is, $`x`$ or $`y`$) on particles $`P_1,P_2,P_3`$, she can only determine whether their results are the same or opposite and also that the eavesdropper will gain no knowledge of what Alice’s and Bob’s actually are, because the eavesdropper will has the probability of 1/2 of making a mistake. In fact, the eavesdropper can not gain any knowledge from Alice even if the measurement bases, she can only obtain the Bob’s measurement bases that is public. By this knowledge, the eavesdropper can not obtain the useful information from Alice and Bob’s sharing key. The eavesdropper can not succeed to obtain the key by the intercept/resend attack defined in . Suppose that the eavesdropper has managed to get a hold of Alice and Bob’s key, she then intercepts Bob’s particle $`P_2`$ and send another particle $`P_2^{}`$ to Bob. In this case, Alice’s and the eavesdropper’s three particles construct a GHZ triplet. Eavesdropper can not obtain the key, because the Alice’s and Bob’s particles are not a GHZ triplet, there are no correlated or anticorrelated result, eavesdropper’s interception will introduce error and can be detected by Alice and Bob. The entanglement attacks is no use in our protocol. To show that, Let us assume that the eavesdropper has been able to entangle an ancilla in state $`|A`$ with the GHZ triplet state that Alice and Bob are using. The state describing the state of the GHZ triplet and the ancilla is $$|\mathrm{\Psi }=\frac{1}{\sqrt{2}}\left(|z+z+z++|zzz\right)|A.$$ (12) By using the $`x`$ and $`y`$ eigenstates and Eq.(7), The eavesdropper get $$\begin{array}{cc}\hfill U|\mathrm{\Psi }=& \frac{1}{2}\left(\right|x+x+|x+|A_1+|xx|x+|A_2\hfill \\ & \\ & +|x+x|x|A_3+|xx+|x|A_4).\hfill \end{array}$$ (13) where $`U`$ denote the unitary transformation. By projecting the above state onto $`|\varphi =\alpha _1|x+x++\alpha _2|xx+\alpha _3|x+x+\alpha _4|xx+`$, the eavesdropper creates the states $$|\mathrm{\Psi }_E=\frac{1}{2}\left(|x+\left(\alpha _1^{}|A_1+\alpha _2^{}|A_2\right)+|x\left(\alpha _3^{}|A_3+\alpha _4^{}|A_4\right)\right).$$ (14) If the eavesdropper can gain the Bob’s qubits, she can obtain the key. However, Eq.(14) shows that the eavesdropper can not obtain Bob’s results. By the similar method, the eavesdropper can not obtain the Bob’s qubits when Alice’s and Bob’s states satify Eqs.(8-10). There is an additional points to notice here. Like all the previous schemes, our scheme itself can also not prevent the men-in-middle attack. The men-in-middle attack may be described as follows: When the legitimate communicator Alice communicates the legitimate communicator Bob, Eve intercepts all qubit sent by Alice, and communicates Bob with impersonating Alice. Finally, Eve obtains two keys $`K_{AE},K_{EB}`$, where $`K_{AE}`$ represents the secret key between Alice and eavesdropper, and $`K_{AE}`$ represents the secret key between Bob and eavesdropper. As a result eavesdropper can easily decrypt the ciphertext sent by Alice or Bob. To prevent the men-in-middle attack, Alice and Bob may verify their identity by using the identity verification technologies . This method need to share a aforehand key between Alice and Bob, the verification proceeding is similar to that presented in . In conclusions, I propose a quantum key distribution protocol based on the GHZ triplet states, the security is warranted by the correlation of the GHZ triplet states. This protocol is unsymmetrical because Alice has two particles and Bob only has one particle from the GHZ triplet. The unsymmetrical characteristic let the proposed protocol have the following merits. 1) The efficiency of the quantum key distribution is excellent, it approaches $`100\%`$. This will be very useful in the practical application. 2) The security is higher than the previous protocol, because Alice’s bases for the quantum key are secret and Alice’s and Bob’s bases are unsymmetrical. I wish to thank Professor G. Guo and Dr. Z. Wang for their useful discussion on the GHZ states. This research was supported by the National Natural Science Foundation of China, Grants No. 69803008. References 1. C. H. Bennett, and G.Brassard, Advances in Cryptology: Proceedings of Crypto 84, August 1984, Springer - Verlag, p. 475. 2. C. H. Bennett, Phys. Rev. Lett., 68, 3121, (1992). 3. A. K.Ekert, Phys. Rev. Lett., 67, 661, (1991). 4. C.H. Bennett, G. Brassard, and N. D. Mermin, Phys. Rev. Lett. 68, 557 (1992). 5. C.A.Fuchs, N.Gisin, R.B.Griffiths, C.S.Niu, and A.Peres, Phys. Rev. A, 56, 1163, (1997). 6. B.A.Slutsky, R.Rao, P.C.Sun, and Y.Fainman, Phys. Rev. A, 57, 2383, (1998). 7. H. E. Brandt, Phys. Rev. A 59, 2665 (1999). 8. C. Niu, and R. Griffiths, Phys. Rev. A, 58, 4377, (1998). 9. C. H. Bennett, F. Bessette, G. Brassard, L. Salvail and J. Smolin, J. Cryptology 5, 3 (1992). 10. J.Breguet, A.Muller, and N.Gisin, Journal of Modern Optics, 41, 2405, (1994). 11. C. Marand and P.D.Townsend, Optics Lett. 20, 1695 (1995). 12. W.T.Buttler, R.J.Hughes, P.G.Kwiat, et. al., Phys. Rev. A 57, 2379 (1998). 13. A.Einstein, B.Podolsky, and N.Rosen, Phys. Rev. 47, 777 (1935). 14. D. Greenberger, M. A. Horne, and A. Zeilinger, in Bell’s Theorem, Quantum theory, and Conceptions of Universe, edited by M.Kaftos (Kluwer Academic, Dordrecht, 1989); D. Greenberger, M. A. Horne, A. Shimony, and A. Zeilinger, Am. J. Phys. 58, 1131 (1990). 15. G. Zeng and W. Zhang, Phys. Rev. A, 61, 2000 (In press).
warning/0001/hep-ph0001193.html
ar5iv
text
# Testing Nonperturbative Ansätze for the QCD Field Strength Correlator ## Acknowledgments The author is indebted to A. Di Giacomo for critical reading the manuscript, useful suggestions and discussions, and E. Meggiolaro for bringing his attention to Ref. and useful discussions. Besides that, he has benefitted from discussions and correspondence with H.G. Dosch, M.G. Schmidt, C. Schubert, Yu.A. Simonov, and K.L. Zarembo. He is also greatful to Prof. A. Di Giacomo and the staff of the Quantum Field Theory Division of the University of Pisa for kind hospitality and INFN for financial support.
warning/0001/hep-ex0001006.html
ar5iv
text
# Search for the Decay 𝐾_𝐿→𝜋⁰⁢𝜇⁺⁢𝜇⁻ ## Abstract We report on a search for the decay $`K_L\pi ^0\mu ^+\mu ^{}`$ carried out as a part of the KTeV experiment at Fermilab. This decay is expected to have a significant $`CP`$ violating contribution and a direct measurement will either support the CKM mechanism for $`CP`$ violation or point to new physics. Two events were observed in the 1997 data with an expected background of $`0.87\pm 0.15`$ events, and we set an upper limit $``$($`K_L\pi ^0\mu ^+\mu ^{}`$)$`<3.8\times 10^{10}`$ at the 90% confidence level. A. Alavi-Harati<sup>12</sup>, I.F. Albuquerque<sup>10</sup>, T. Alexopoulos<sup>12</sup>, M. Arenton<sup>11</sup>, K. Arisaka<sup>2</sup>, S. Averitte<sup>10</sup>, A.R. Barker<sup>5</sup>, L. Bellantoni<sup>7,†</sup>, A. Bellavance<sup>9</sup>, J. Belz<sup>10</sup>, R. Ben-David<sup>7</sup>, D.R. Bergman<sup>10</sup>, E. Blucher<sup>4</sup>, G.J. Bock<sup>7</sup>, C. Bown<sup>4</sup>, S. Bright<sup>4</sup>, E. Cheu<sup>1</sup>, S. Childress<sup>7</sup>, R. Coleman<sup>7</sup>, M.D. Corcoran<sup>9</sup>, G. Corti<sup>11</sup>, B. Cox<sup>11</sup>, M.B. Crisler<sup>7</sup>, A.R. Erwin<sup>12</sup>, R. Ford<sup>7</sup>, A. Glazov<sup>4</sup>, A. Golossanov<sup>11</sup>, G. Graham<sup>4</sup>, J. Graham<sup>4</sup>, K. Hagan<sup>11</sup>, E. Halkiadakis<sup>10</sup>, K. Hanagaki<sup>8</sup>, M. Hazumi<sup>8</sup>, S. Hidaka<sup>8</sup>, Y.B. Hsiung<sup>7</sup>, V. Jejer<sup>11</sup>, J. Jennings<sup>2</sup>, D.A. Jensen<sup>7</sup>, R. Kessler<sup>4</sup>, H.G.E. Kobrak<sup>3</sup>, J. LaDue<sup>5</sup>, A. Lath<sup>10</sup>, A. Ledovskoy<sup>11</sup>, P.L. McBride<sup>7</sup>, A.P. McManus<sup>11</sup>, P. Mikelsons<sup>5</sup>, E. Monnier<sup>4,∗</sup>, T. Nakaya<sup>7</sup>, K.S. Nelson<sup>11</sup>, H. Nguyen<sup>7</sup>, V. O’Dell<sup>7</sup>, M. Pang<sup>7</sup>, R. Pordes<sup>7</sup>, V. Prasad<sup>4</sup>, C. Qiao<sup>4</sup>, B. Quinn<sup>4</sup>, E.J. Ramberg<sup>7</sup>, R.E. Ray<sup>7</sup>, A. Roodman<sup>4</sup>, M. Sadamoto<sup>8</sup>, S. Schnetzer<sup>10</sup>, K. Senyo<sup>8</sup>, P. Shanahan<sup>7</sup>, P.S. Shawhan<sup>4</sup>, W. Slater<sup>2</sup>, N. Solomey<sup>4</sup>, S.V. Somalwar<sup>10</sup>, R.L. Stone<sup>10</sup>, I. Suzuki<sup>8</sup>, E.C. Swallow<sup>4,6</sup>, R.A. Swanson<sup>3</sup>, S.A. Taegar<sup>1</sup>, R.J. Tesarek<sup>10</sup>, G.B. Thomson<sup>10</sup>, P.A. Toale<sup>5</sup>, A. Tripathi<sup>2</sup>, R. Tschirhart<sup>7</sup>, Y.W. Wah<sup>4</sup>, J. Wang<sup>1</sup>, H.B. White<sup>7</sup>, J. Whitmore<sup>7</sup>, B. Winstein<sup>4</sup>, R. Winston<sup>4</sup>, T. Yamanaka<sup>8</sup>, E.D. Zimmerman<sup>4</sup> <sup>1</sup> University of Arizona, Tucson, AZ 85721 <sup>2</sup> University of California at Los Angeles, Los Angeles, CA 90095 <sup>3</sup> University of California at San Diego, La Jolla, CA 92093 <sup>4</sup> The Enrico Fermi Institute, The University of Chicago, Chicago, IL 60637 <sup>5</sup> University of Colorado, Boulder, CO 80309 <sup>6</sup> Elmhurst College, Elmhurst, IL 60126 <sup>7</sup> Fermi National Accelerator Laboratory, Batavia, IL 60510 <sup>8</sup> Osaka University, Toyonaka, Osaka 560 Japan <sup>9</sup> Rice University, Houston, TX 77005 <sup>10</sup> Rutgers University, Piscataway, NJ 08855 <sup>11</sup> The Department of Physics and Institute of Nuclear and Particle Physics, University of Virginia, Charlottesville, VA 22901 <sup>12</sup> University of Wisconsin, Madison, WI 53706 On leave from C.P.P. Marseille/C.N.R.S., France To whom correspondence should be addressed. The decays $`K_L\pi ^0l\overline{l}`$ are interesting decays for the study of $`CP`$ violation and can be used to search for new physics. There are three expected contributions to the amplitude: a $`CP`$ conserving contribution which proceeds through the $`\pi ^0\gamma ^{}\gamma ^{}`$ intermediate state, an indirectly $`CP`$ violating contribution from $`K_1\pi ^0l\overline{l}`$, and a directly $`CP`$ violating contribution from electroweak penguin and $`W`$ box diagrams . Branching ratio predictions in theories containing exotic (e.g., SUSY) particles that contribute to the penguin amplitudes are significantly higher . The sizes of the three contributions depend on the flavor of the final state lepton. The greatest theoretical interest is in the $`K_L\pi ^0\nu \overline{\nu }`$ case, where the direct $`CP`$ violating amplitude dominates and a theoretically clean measurement of the Wolfenstein parameter $`\eta `$ should be possible. However, measuring a final state with two neutrinos and a single pion is experimentally challenging, and the current experimental limit remains four orders of magnitude above the Standard Model expectation of $`3\times 10^{11}`$. In contrast, $`K_L\pi ^0e^+e^{}`$ and $`K_L\pi ^0\mu ^+\mu ^{}`$ are comparatively straightforward to detect, although all three amplitudes are present in these modes. This letter presents a new limit on $``$($`K_L\pi ^0\mu ^+\mu ^{}`$ ); the existing limit is $`5.1\times 10^9`$ at the 90% C.L. For the data taken by KTeV in 1997 and discussed here, a single event observed in the muon mode would correspond to a branching ratio of $`7\times 10^{11}`$, which approaches the Standard Model expectation of $``$($`K_L\pi ^0\mu ^+\mu ^{}`$)$`(0.441.00)\times 10^{11}`$. Figure 1 shows a plan view of the KTeV detector, which has been described elsewhere . An 800 GeV proton beam, with typically $`3.5\times 10^{12}`$ protons per 19 s Fermilab Tevatron spill every minute, was targeted at a vertical angle of 4.8 mrad on a 1.1 interaction length (30 cm) BeO target. Photons were converted by 76 mm of lead immediately downstream of the target. Charged particles were then removed with magnetic sweeping. Collimators defined two 0.25 $`\mu `$sr beams that entered the KTeV apparatus 94 m downstream of the target. The 65 m vacuum ($`10^6`$ Torr) decay region extended to the first drift chamber. The spectrometer consisted of a dipole magnet surrounded by four ($`1.28\times 1.28`$ m<sup>2</sup> to $`1.77\times 1.77`$ m<sup>2</sup>) drift chambers with $``$100$`\mu `$m position resolution in both horizontal and vertical directions. Helium filled bags occupied the spaces between the drift chambers; the magnetic field imparted a $`\pm `$205 MeV/c horizontal momentum kick. The spectrometer had a momentum resolution of $`\sigma (P)/P=0.38\%0.016\%P`$, where $`P`$ is in GeV/c. The electromagnetic calorimeter consisted of 3100 pure CsI crystals. Each crystal was 50 cm (27 radiation lengths, 1.4 interaction lengths) long. Crystals in the central $`1.2\times 1.2`$ m<sup>2</sup> section of the calorimeter had a cross-sectional area of $`2.5\times 2.5`$ cm<sup>2</sup>, and those in the outer region (out to $`1.9\times 1.9`$ m<sup>2</sup>) had a $`5\times 5`$ cm<sup>2</sup> area. The calorimeter’s energy resolution for photons was $`\sigma (E)/E=0.45\%2\%/\sqrt{E}`$, where $`E`$ is in\̇mbox{\\,GeV}, and its position resolution was $``$1 mm. Nine photon veto assemblies (lead scintillator sandwiches) detected particles leaving the fiducial volume. Two scintillator hodoscopes in front of the calorimeter were used to trigger on charged particles. The hodoscopes and the calorimeter had two holes ($`15\times 15`$ cm at the calorimeter) to let the neutral beams pass through without interaction. The muon filter, located behind the calorimeter, was constructed of a 10 cm thick lead wall followed by three steel walls totaling 511 cm thickness. Scintillator planes with 15 cm segmentation in both horizontal and vertical directions (MU3) were located after the third steel wall. The segmentation was comparable to the multiple scattering angle of 10 GeV muons at MU3. Pion punch-through probabilities, including decays downstream of the calorimeter, were taken as a function of momentum from $`K_L\pi ^\pm e^{}\nu `$ data and are on the order of $`2\times 10^3`$. The data acquisition system reconstructed events online, and the results were used to filter the data. The signature we searched for is two tracks from oppositely charged particles with a common vertex that deposit little energy in the calorimeter and created two hits at MU3 from these muons. The $`\pi ^0`$ creates two electromagnetic showers in the calorimeter with $`m_{\gamma \gamma }=m_{\pi ^0}`$ and which are unassociated to tracks. There are three important backgrounds. The first is $`K_L\pi ^+\pi ^{}\pi ^0`$ where both $`\pi ^\pm `$ either decay upstream of the calorimeter (decay in flight) or punch through to MU3. The second is $`K_L\pi ^\pm \mu ^{}\nu `$ with one decay in flight or punch-through and accidentally coincident calorimeter activity that appears as a $`\pi ^0`$. The third and largest background is the radiative muonic Dalitz decay $`K_L\mu ^+\mu ^{}\gamma \gamma `$ when $`m_{\gamma \gamma }=m_{\pi ^0}`$. Because of the low expected branching ratio, $`K_L\pi ^0\pi ^\pm \mu ^{}\nu `$ is not a large background. Two triggers were used for this analysis. To determine the number of $`K_L`$ decays in the data, we identified $`K_L\pi ^+\pi ^{}\pi ^0`$ decays in a minimum bias trigger. This trigger required hits in the trigger hodoscopes and the drift chambers which were consistent with two coincident charged particles passing through the detector. Events with a reconstructed vertex from oppositely charged tracks were recorded with a prescale factor of 500:1. For the signal trigger further requirements were made. Two or more hits in MU3 were required, and activity in the photon veto counters rejected events. The trigger system counted the number of calorimeter clusters over $`1`$ GeV in a narrow (20nsec) time gate; for the signal mode, at least one such cluster was required. We also required that the calorimeter energy reconstructed online and associated with each track be less than 5 GeV. The signal trigger was not prescaled. Muons passing through the calorimeter typically deposit $``$400 MeV, and so in searching offline for the signal we required that the clusters associated to the tracks had less than 1 GeV of energy. Track momenta were required to be greater than 10 GeV/c to ensure that they penetrated to MU3 and less than 100 GeV/c to ensure that the momentum was well measured. The two-muon system was required to have a mass less than 350 MeV/c<sup>2</sup> to reduce backgrounds from $`K_L\pi ^\pm \mu ^{}\nu `$. We required two non-adjacent hits in both views of MU3. We did not compare the MU3 hit positions to the extrapolation of the tracks to MU3, because the major backgrounds passed this requirement as well as the signal did. To suppress $`\pi ^\pm `$ decays in flight, we required that the reconstructed vertex occurred in the beam volume of the decay region and had a $`\chi ^2`$ of 10 for 1 d.o.f. or less, and that the track segments upstream and downstream of the spectrometer magnet passed within 1 mm (94% signal acceptance) of each other at the bend plane of the magnet. The mass of the two unassociated clusters under the hypothesis that they were produced by photons from the decay vertex was required to be between 135$`\pm `$6 MeV/c<sup>2</sup> ($`\pm 2.5\sigma `$). These clusters both had to have been found by the trigger cluster counter. A number of kinematic criteria were studied to suppress background from $`K_L\mu ^+\mu ^{}\gamma \gamma `$ decays. The kinematics of this decay are very different from the analogous $`K_Le^+e^{}\gamma \gamma `$ background to $`K_L\pi ^0e^+e^{}`$, and the branching ratio is lower. Consequently the methods which are effective in the $`e^\pm `$ case are less effective in the $`\mu ^\pm `$ case and were not applied. It is possible to suppress backgrounds from $`K_L\pi ^+\pi ^{}\pi ^0`$, and $`K_L\pi ^\pm \mu ^{}\nu `$ by using $$\mathrm{R}_{}^{\mu \mu }\frac{(m_K^2m_{\mu \mu }^2m_{\pi ^0}^2)^24m_{\mu \mu }^2m_{\pi ^0}^24m_K^2p_{\mu \mu }^2}{p_{\mu \mu }^2+m_{\mu \mu }^2}$$ (1) where $`m_K`$ is the kaon mass, $`m_{\mu \mu }`$ the mass of the two-muon system, $`m_{\pi ^0}`$ the $`\pi ^0`$ mass, and $`p_{\mu \mu }`$ is the two-muon systems’s momentum perpendicular to the kaon momentum. This quantity is proportional to the $`\pi ^0`$ momentum squared in the $`K_L`$ flight direction in the frame colinear with the $`K_L`$ but where the $`\mu ^+\mu ^{}`$ pair has no longitudinal momentum. We required $`\mathrm{R}_{}^{\mu \mu }`$ to lie between -0.01 and 0.10 $`GeV^2/c^4`$. This cut keeps 89.2% of the signal and rejects 73% of $`K_L\pi ^\pm \mu ^{}\nu `$ decays with coincident photons and 95% of the $`K_L\pi ^+\pi ^{}\pi ^0`$ decays. To ensure that we observed all the products of a $`K_L`$ decay, we required that the total squared momentum transverse to the $`K_L`$ flight direction ($`P_{}^2`$) be less than 100 $`MeV^2/c^2`$, and that the reconstructed mass ($`m`$) of the $`K_L`$ be between 492 and 504 MeV/c<sup>2</sup>. With these requirements, which were selected by examining Monte Carlo simulation results and data outside the signal region before examining the data for $`K_L\pi ^0\mu ^+\mu ^{}`$ candidates, the overall acceptance for the signal was 5.0%. The simulation distributed the products of the decay uniformly in phase space. Figure 2 shows the $`P_{}`$<sup>2</sup> vs. $`m`$ distributions for the data, and Fig. 3 shows the mass distribution after the $`P_{}^2`$ requirement for the data and the backgrounds as estimated from the simulation. The correspondence between the data and the simulation is good, and is also good in the distributions of $`m_{\mu \mu }`$, $`P_{}^2`$, track momentum, vertex position and (for $`K_L\pi ^+\pi ^{}\pi ^0`$) of $`m_{\gamma \gamma }`$. The background levels in the signal region are given in Table I; they are calculated from the simulations, published branching ratios, and the number of $`K_L`$ decays in the data sample. Although we have observed $`K_L\mu ^+\mu ^{}\gamma \gamma `$ in this data, this background was more precisely estimated from QED and the measured $``$($`K_L\mu ^+\mu ^{}\gamma `$); a value of $``$($`K_L\mu ^+\mu ^{}\gamma \gamma `$)$`=(9.1\pm 0.8)\times 10^9`$ with an $`m_{\gamma \gamma }>1`$ MeV/c<sup>2</sup> cutoff was used. All Monte Carlo samples were over 10 times the data sample. The background from $`K_L\pi ^+\pi ^{}`$ \+ $`2\gamma `$ (Acc) and $`K_L\pi ^+\pi ^{}\gamma +\gamma `$ (Acc) was negligible. To normalize any possible signal’s branching ratio, we identified $`K_L\pi ^+\pi ^{}\pi ^0`$ decays in as similar a manner to the $`K_L\pi ^0\mu ^+\mu ^{}`$ identification as possible. Apart from the trigger differences, the calorimeter energy requirement for clusters associated to tracks was changed to less than 0.9 times the momentum measured with the spectrometer; the MU3 and $`\mathrm{R}_{}^{\mu \mu }`$ requirements were removed, and $`m_{\pi ^\pm }`$ rather than $`m_{\mu ^\pm }`$ was used in calculating kinematic quantities. The acceptance for the normalization mode was 8.1%. There were $`(268\pm 0.4_{\mathrm{S}TAT}\pm 0.4_{\mathrm{M}C}\pm 4.3_{\mathrm{B}R})\times 10^9`$$`K_L`$ decays between 90 and 160 m from the target with $`K_L`$ momentum between 20 and 220 GeV/c. In calculating the number of $`K_L`$ decays in the data and the acceptance for $`K_L\pi ^0\mu ^+\mu ^{}`$ relative to the acceptance for $`K_L\pi ^+\pi ^{}\pi ^0`$, we allowed for the uncertainties summarized in Table II. We calculated the $`K_L`$ flux using $`K_L\pi ^\pm \mu ^{}\nu `$ rather than $`K_L\pi ^+\pi ^{}\pi ^0`$ decays and attributed the difference of 4.20% to the quality of our simulation of muons in the detector. We varied the scale and resolution of the calorimeter and spectrometer in the simulation to conservatively cover the range of variations seen in the data. Apart from uncertainties in published branching ratios, other sources of uncertainty were small. From Figs. 2 and 3, two events exist in the signal region for the data. Sidebands in both $`m`$ and $`P_{}`$<sup>2</sup> show correspondence between data and background predictions as given in Table III. With the above acceptance and $`K_L`$ flux, and allowing for a background level of $`0.87\pm 0.15`$ events, we set an upper limit $``$($`K_L\pi ^0\mu ^+\mu ^{}`$)$`<3.8\times 10^{10}`$ at the 90% confidence level. This limit is approximately one order of magnitude more stringent than the previous limit. If we assume that the only contribution to the branching ratio is from direct $`CP`$ violation, we may conclude that $`|\eta |<7`$ at the 90% confidence level. In comparison to the search for $`K_L\pi ^0e^+e^{}`$, $`K_L\pi ^0\mu ^+\mu ^{}`$ searches will have better single event sensitivity for any given sample of $`K_L`$ decays because the level of irreducible $`K_Ll^+l^{}\gamma \gamma `$ background is less. While not yet as sensitive as B decays, where a recent indirect global analysis finds $`\eta `$ to be below 1, it is valuable to test if the same parameterization is valid for both B and K decays. We gratefully acknowledge the support and effort of the Fermilab staff and the technical staffs of the participating institutions for their vital contributions. This work was supported in part by the U.S. Department of Energy, The National Science Foundation and The Ministry of Education and Science of Japan. In addition, A.R.B., E.B. and S.V.S. acknowledge support from the NYI program of the NSF; A.R.B. and E.B. from the Alfred P. Sloan Foundation; E.B. from the OJI program of the DOE; K.H., K.S., T.N. and M.S. from the Japan Society for the Promotion of Science.
warning/0001/hep-ph0001074.html
ar5iv
text
# Discovery and Identification of Extra Gauge Bosons in 𝑒⁺⁢𝑒⁻→𝜈⁢𝜈̄⁢𝛾 ## I Introduction Extra gauge bosons, both charged ($`W^{}`$) and/or neutral ($`Z^{}`$), arise in many models of physics beyond the Standard Model (SM) . Examples include extended gauge theories such as grand unified theories and Left-Right symmetric models along with the corresponding supersymmetric models, and other models such as those with finite size extra dimensions . To elucidate what physics lies beyond the Standard Model it is necessary to search for manifestations of that new physics with respect to the predicted particle content, both fermions and extra gauge bosons. Such searches are a feature of ongoing collider experiments and the focus of future experiments. The discovery of new particles would provide definitive evidence for physics beyond the Standard Model and, in particular, the discovery of new gauge bosons would indicate that the standard model gauge group was in need of extension. There is a considerable literature on $`Z^{}`$ searches. In this paper we concentrate on $`W^{}`$ searches, for which much less work has been done. Limits have been placed on the existence of new gauge bosons through indirect searches based on the deviations from the SM they would produce in precision electroweak measurements. For instance, indirect limits from $`\mu `$-decay constrain the LRM $`W^{}`$ to $`M_{W_{LR}^{}}550`$ GeV . A more severe constraint arises from $`K_LK_S`$ mass-splitting which gives $`M_{W_{LR}^{}}1.6`$ TeV . In obtaining the above limits, it was assumed that the coupling constants of the two $`SU(2)`$ gauge groups are equal. New gauge boson searches at hadron colliders consider their direct production via the Drell-Yan process and their subsequent decay to lepton pairs. For $`W^{}`$ bosons, decays to hadronic jets are sometimes also considered. The present bounds on neutral gauge bosons, $`Z^{}`$’s, from the CDF and D0 collaborations at the Tevatron $`p\overline{p}`$ collider at Fermilab are $`M_Z^{}>590690`$ GeV with the exact value depending on the specific model . For $`W^{}`$’s the limits are $`M_W^{}>300720`$ GeV; again the limits depend on the details of the model . The search reach is expected to increase by $`300`$ GeV with 1 fb<sup>-1</sup> of luminosity . The Large Hadron Collider is expected to be able to discover $`Z^{}`$’s up to masses of 4-5 TeV and $`W^{}`$’s up to masses of $`5.9`$ TeV . The $`W^{}`$ limits assume SM strength couplings and decay into a light stable neutrino which is registered in the detector as missing $`E_T`$. They can be seriously degraded by loosening the assumptions in the model. In addition, one can place limits on new gauge bosons by looking for deviations from SM expectations for observables measured at $`ep`$ and $`e^+e^{}`$ colliders. Searches for new gauge bosons at $`e^+e^{}`$ colliders are kinematically limited by the available center-of-mass energy so that one searches for indirect effects of extra gauge bosons in cross sections and asymmetries for $`\sqrt{s}<M_V^{}`$. There is a considerable body of work on $`Z^{}`$ searches at $`e^+e^{}`$ colliders and, although the discovery limits are very model dependent, they lie in the general range of 2-5 TeV for $`\sqrt{s}=500`$ GeV with 50 fb<sup>-1</sup> luminosity . In contrast to the $`Z^{}`$ case, there are virtually no studies of indirect searches for $`W^{}`$ bosons at $`e^+e^{}`$ colliders. Recently, Hewett suggested that the reaction $`e^+e^{}\nu \overline{\nu }\gamma `$ would be sensitive to $`W^{}`$’s with masses greater than $`\sqrt{s}`$ . In the Standard Model, this process proceeds through $`s`$-channel $`Z`$ and $`t`$-channel $`W`$ exchange with the photon being radiated from every possible charged particle. In extended gauge models the process is modified by both $`s`$-channel $`Z^{}`$ and $`t`$-channel $`W^{}`$ exchange. In this paper, we examine this process for various extended electroweak models. The first model we consider is the Left-Right symmetric model based on the gauge group $`SU(3)_C\times SU(2)_L\times SU(2)_R\times U(1)_{BL}`$ which has right-handed charged currents. The second model we consider is the Un-Unified model which is based on the gauge group $`SU(2)_q\times SU(2)_l\times U(1)_Y`$ where the quarks and leptons each transform under their own $`SU(2)`$. The final type of model, which has received considerable interest lately, contains the Kaluza-Klein excitations of the SM gauge bosons which are a possible consequence of theories with large extra dimensions . The models under consideration are described in more detail in Section II. Additionally, we study discovery limits for various combinations of $`W^{}`$ and $`Z^{}`$ bosons with SM couplings. Although these are not realistic models, they have been adopted as benchmarks to compare the discovery reach of different processes. We will find that, while the process $`e^+e^{}\nu \overline{\nu }\gamma `$ can indeed extend the discovery reach for $`W^{}`$’s significantly beyond $`\sqrt{s}`$, with the exact limit depending on the specific model, it is not in general competitive with limits obtainable at the LHC. However, if extra gauge bosons are discovered which are not overly massive, the process considered here could be used to measure their couplings. This would be crucial for determining the origins of the $`Z^{}`$ or $`W^{}`$. As such, it would play an important complementary role to the LHC studies. In the next section we review the relevant details of the various models that we use in our calculations. In Section III, we describe the details of our calculations. The resulting $`W^{}`$ discovery limits and projected sensitivities for $`W^{}`$ couplings and $`Z^{}\nu \overline{\nu }`$ couplings are given in Section IV. We conclude with some final comments. ## II Models In this section, we describe the models considered in our investigation. The so-called Sequential Standard Model (SSM) includes additional weak gauge bosons of higher mass, with SM couplings. This is a rather arbitrary scenario which we include only as a benchmark. Since our emphasis here is on extra $`W`$’s, we consider a SSM with a $`W^{}`$ only, which we refer to as SSM($`W^{}`$), and a SSM with both $`W^{}`$ and $`Z^{}`$, denoted by SSM($`W^{}+Z^{}`$). In the latter, we will take $`M_Z^{}=M_W^{}`$ for simplicity. The general Left-Right symmetric model (LRM) is based on the extended electroweak gauge group $`SU(2)_L\times SU(2)_R\times U(1)_{BL}`$. Left-handed fermion fields transform as doublets under $`SU(2)_L`$ and as singlets under $`SU(2)_R`$. The reverse is true for right-handed fermions. A right-handed neutrino is included in the fermion content. The model is parametrized by the ratio of the coupling constants of the two $`SU(2)`$ gauge groups, which we denote as $`\kappa =g_R/g_L`$. This parameter is allowed to vary here in the approximate range $`0.55\kappa 2.0`$ . The lower bound on $`\kappa `$ arises from the condition $`\mathrm{sin}^2\theta _W\frac{\kappa ^2}{1+\kappa ^2}`$ (or, equivalently, $`\kappa ^2\mathrm{tan}^2\theta _W`$), which expresses the positivity of a ratio of squared couplings. In principle, $`\kappa `$ is restricted to be less than 1 based on symmetry breaking scenarios and coupling constant evolution arguments. However, it is conceivable that this bound may be violated in some Grand Unified Theory so we take a phenomenological approach and loosen this upper bound . Additionally, a parameter, $`\rho `$, describes the Higgs content of the model. If only Higgs doublets are used to break the gauge symmetry to $`U(1)_{em}`$, $`\rho `$ is 1. For Higgs triplets, $`\rho `$ is 2. A combination of doublets and triplets leads to an intermediate value of $`\rho `$ between 1 and 2 . We will use $`\rho =1`$, corresponding to Higgs doublets. In the LRM, there is a relationship between the $`Z^{}`$ and $`W^{}`$ masses, as follows: $$\frac{M_Z^{}^2}{M_W^{}^2}=\frac{\rho \kappa ^2}{\kappa ^2\mathrm{tan}^2\theta _W}.$$ (1) The couplings of the extra gauge bosons relevant to our calculation can be read from the following parts of the Lagrangian. $`_{LR}`$ $`=`$ $`{\displaystyle \frac{e\kappa }{\sqrt{2}s_W}}W_\mu ^+\overline{\nu }_R\gamma ^\mu e_R+{\displaystyle \frac{e}{2s_Wc_W^2\sqrt{\kappa ^2t_W^2}}}Z_\mu ^{}[\overline{l}\gamma ^\mu (1\gamma _5)s_W^2(T_{3L}Q_{em})l`$ (3) $`+\overline{l}\gamma ^\mu (1+\gamma _5)(\kappa ^2c_W^2T_{3R}s_W^2Q_{em})l]+h.c.`$ where $`e_R=\frac{1}{2}(1+\gamma _5)e`$ denotes a right-handed electron field. Note that we neglect two angles, usually denoted as $`\xi `$ and $`\zeta `$, which parametrize the $`ZZ^{}`$ and $`WW^{}`$ mixings, respectively. Limits on these angles are rather severe so this is justified . Neglect of these angles implies SM couplings for the $`Z`$ and $`W`$. Additionally, we assume light Dirac-type neutrinos. The Un-Unified model (UUM) employs the alternative electroweak gauge symmetry $`SU(2)_q\times SU(2)_l\times U(1)_Y`$ with left-handed quarks and leptons transforming as doublets under their respective $`SU(2)`$ groups. All the right-handed fields transform as singlets under both $`SU(2)`$ groups. The UUM may be parametrized by an angle $`\varphi `$, which represents the mixing of the charged gauge bosons of the two $`SU(2)`$ groups, and by a ratio $`x=(u/v)^2`$, where $`u`$ and $`v`$ are the vacuum expectation values of the scalar multiplets which break the symmetry to $`U(1)_{em}`$. The existing constraint on $`\varphi `$ is $`0.24\mathrm{sin}\varphi 0.99`$, based on the validity of perturbation theory. For $`x/\mathrm{sin}^2\varphi 1`$, the $`Z^{}`$ mass is approximately equal to that of the $`W^{}`$ and the parameter $`x`$ may be replaced by $`M_W^{}`$. The lepton couplings of interest to us here arise from the following part of the Lagrangian. $`_{UU}`$ $`=`$ $`{\displaystyle \frac{e}{2s_W}}{\displaystyle \frac{s_\varphi }{c_\varphi }}\left[\sqrt{2}W_\mu ^+\overline{\nu }\gamma ^\mu l_L+Z_\mu ^{}(\overline{\nu }\gamma ^\mu \nu _L\overline{l}\gamma ^\mu l_L)\right]+h.c.`$ (4) As expected, the fermion couplings to the additional gauge bosons are all left-handed in the UUM. Additional fermions must also be included in order to cancel anomalies. This is rather difficult to do without generating flavour changing neutral currents and some considerations of this problem lead to rather high lower bounds on the $`Z^{}`$ mass of about 1.4 TeV . However, lower $`Z^{}`$ and $`W^{}`$ masses may be allowed in other scenarios; hence we take a phenomenological approach in this investigation. Finally, we consider the consequences of models which have been of considerable interest lately, those containing large extra dimensions . In particular, we consider an extension of the SM to 5-dimensions (5DSM) . The presence of an extra dimension of size $`R`$ TeV<sup>-1</sup> may imply an infinite tower of Kaluza-Klein (KK) excitations of the SM gauge bosons. The mass of the excitations is associated with the compactification scale of the extra dimension as $`nM_c`$ ($`n=1,\mathrm{},\mathrm{}`$), where $`M_c=1/R`$. The properties of and relationships among electroweak observables are modified by the presence of these KK towers. We treat this possibility in a manner similar to the other models described above; that is, we include in our process the exchange of a $`W^{}`$ and $`Z^{}`$ corresponding to the first KK excitations. The model can be parametrized by an angle $`\beta `$ which is correlated with the properties of its Higgs sector, which includes two doublets; for $`\mathrm{sin}\beta s_\beta =0`$, the SM Higgs may propagate in all 5-dimensions (the bulk) while for $`s_\beta =1`$, it is confined to the 4-dimensional boundary. In terms of this parameter, the physical masses of the lightest electroweak gauge bosons (corresponding to the experimentally measured masses) are given, to first order in $`M_W^2/M_c^2`$, as $`M_W^{(ph)2}`$ $`=`$ $`M_W^2\left[1s_\beta ^4{\displaystyle \frac{\pi ^2}{3}}{\displaystyle \frac{M_W^2}{M_c^2}}\right]`$ (5) $`M_Z^{(ph)2}`$ $`=`$ $`M_Z^2\left[1s_\beta ^4{\displaystyle \frac{\pi ^2}{3}}{\displaystyle \frac{M_Z^2}{M_c^2}}\right]`$ (6) where $`M_W^2=g^2v^2/2`$, as usual.The gauge couplings of the physical $`W`$ and $`Z`$ are also modified by a term of order $`M_V^2/M_c^2`$. Global analyses of electroweak parameters put a lower limit on $`M_c`$ of about 2.5 TeV so this is a very small effect. We will therefore neglect it and thus eliminate $`s_\beta `$ as a parameter. On the other hand, the fermion coupling of the first KK excitations, $`W^{}`$ and $`Z^{}`$, each of mass $`M_V^{}=M_c`$, is enhanced by a factor of $`\sqrt{2}`$. Hence our consideration of the 5DSM amounts to including a $`W^{}`$ and a $`Z^{}`$, of equal mass, each coupling as in the SM apart from an extra factor of $`\sqrt{2}`$. ## III Calculation The process under consideration is $$e^{}(p_{})+e^+(p_+)\gamma (k)+\nu (q_{})+\overline{\nu }(q_+).$$ (7) The relevant Feynman diagrams are given in Fig. 1. The kinematic observables of interest are the photon’s energy, $`E_\gamma `$, and its angle relative to the incident electron, $`\theta _\gamma `$, both defined in the $`e^+e^{}`$ center-of-mass frame. The invariant mass of the $`\nu \overline{\nu }`$ pair, $`M_{\nu \overline{\nu }}`$, and $`E_\gamma `$ are related via $$E_\gamma =\frac{\sqrt{s}}{2}\left(1\frac{M_{\nu \overline{\nu }}^2}{s}\right),$$ (8) where $`s=(p_++p_{})^2`$. Let $``$ denote the sum of the amplitudes shown in Fig. 1, over a given number of $`Z^{}`$’s and $`W^{}`$’s. The doubly differential cross section is related to $`||^2`$ via $$\frac{d\sigma }{dE_\gamma d\mathrm{cos}\theta _\gamma }=\frac{E_\gamma }{2s}\frac{1}{(4\pi )^4}_0^\pi 𝑑\theta \mathrm{sin}\theta _0^{2\pi }𝑑\phi ||^2,$$ (9) where $`\theta `$ and $`\phi `$ are the polar and azimuthal angles, respectively, of $`q_+`$ in a frame where $`q_+`$ and $`q_{}`$ are back-to-back. The explicit momentum parametrizations are given in the Appendix. Two approaches to determining $`||^2`$ are possible. One can determine $``$ analytically, using spinor techniques for instance, then square it numerically or one can find $`||^2`$ analytically. We have followed both approaches, which provides an independent check. Obtaining $`||^2`$ analytically has been done both via the trace method, using the symbolic manipulation program FORM , and by squaring the helicity amplitudes and summing over the final state helicities. The latter approach leads to a rather compact result which we present below. In order to present $`||^2`$, we define the following kinematic variables. We follow the notation of , where the SM contribution for this process was calculated: $$\begin{array}{ccccc}& \hfill s=& (p_++p_{})^2,\hfill & \hfill s^{}=& (q_++q_{})^2,\hfill \\ & \hfill t=& (p_+q_+)^2,\hfill & \hfill t^{}=& (p_{}q_{})^2,\hfill \\ & \hfill u=& (p_+q_{})^2,\hfill & \hfill u^{}=& (p_{}q_+)^2,\hfill \\ & \hfill k_\pm =& 2p_\pm k,\hfill & \hfill k_\pm ^{}=& 2q_\pm k,\hfill \\ \hfill Z_i=s^{}M_{Z_i^2}+iM_{Z_i}\mathrm{\Gamma }_{Z_i},& \hfill W_i=& tM_{W_i^2},\hfill & \hfill W_i^{}=& t^{}M_{W_i^2}.\hfill \end{array}$$ (10) The decay width of the extra neutral gauge boson, $`\mathrm{\Gamma }_{Z_i}`$, into fermion- antifermion pairs is calculated in each of the models we consider. We include the one-loop QED, three-loop QCD and $`O(M_t^2/M_Z^{}^2)`$ corrections, although their effect on the cross section is negligible. In the following, we denote generalized couplings as may be inferred from the vertices $`Z_if\overline{f}`$ $`=`$ $`{\displaystyle \frac{ig}{2c_W}}\gamma ^\mu \left({\displaystyle \frac{1\gamma _5}{2}}a_{Z_i}^f+{\displaystyle \frac{1+\gamma _5}{2}}b_{Z_i}^f\right)`$ (11) $`W_il\nu `$ $`=`$ $`{\displaystyle \frac{ig}{\sqrt{2}}}\gamma ^\mu \left({\displaystyle \frac{1\gamma _5}{2}}a_{W_i}+{\displaystyle \frac{1+\gamma _5}{2}}b_{W_i}\right).`$ (12) Thus, in the SM, $`a_{Z_1}^e=2s_W^21`$, $`b_{Z_1}^e=2s_W^2`$, $`a_{Z_1}^\nu =1`$, $`b_{Z_1}^\nu =0`$, $`a_{W_1}=1`$, and $`b_{W_1}=0`$. It is only necessary to present the unpolarized squared amplitude as the individual polarized contributions may be inferred from the coupling structure. The spin-averaged unpolarized $`||^2`$ is given by: $`||_{\mathrm{unp}}^2`$ $`=`$ $`{\displaystyle \frac{(4\pi )^3\alpha ^3}{8s_W^4k_+k_{}}}\{{\displaystyle \frac{3s^{}}{c_W^4}}{\displaystyle \underset{\stackrel{i=1,nz}{j=i,nz}}{}}Z_{ij}[(a_{Z_i}^ea_{Z_j}^ea_{Z_i}^\nu a_{Z_j}^\nu +b_{Z_i}^eb_{Z_j}^eb_{Z_i}^\nu b_{Z_j}^\nu )(u^2+u^2)`$ (17) $`+(a_{Z_i}^ea_{Z_j}^eb_{Z_i}^\nu b_{Z_j}^\nu +b_{Z_i}^eb_{Z_j}^ea_{Z_i}^\nu a_{Z_j}^\nu )(t^2+t^2)]`$ $`+{\displaystyle \frac{4}{s^{}}}{\displaystyle \underset{\stackrel{i=1,nw}{j=i,nw}}{}}W_{ij}[(a_{W_i}^2a_{W_j}^2+b_{W_i}^2b_{W_j}^2)(u^2+u^2)+2a_{W_i}a_{W_j}b_{W_i}b_{W_j}(s^2+s^2)]`$ $`+{\displaystyle \frac{4}{c_W^2}}{\displaystyle \underset{\stackrel{i=1,nw}{j=1,nz}}{}}[\left(WZ\right)_{ij}(u^2a_{W_i}^2a_{Z_j}^ea_{Z_j}^\nu +u^2b_{W_i}^2b_{Z_j}^eb_{Z_j}^\nu )`$ $`+\left(WZ\right)_{ij}^{}(u^2a_{W_i}^2a_{Z_j}^ea_{Z_j}^\nu +u^2b_{W_i}^2b_{Z_j}^eb_{Z_j}^\nu )]\},`$ where $$\begin{array}{cccccc}\hfill Z_{ij}& =& \mathrm{Re}\left(\frac{2\delta _{ij}}{Z_iZ_j^{}}\right),\hfill & \hfill W_{ij}& =& (2\delta _{ij})\mathrm{Re}(F_{W_i}F_{W_j}^{}),\left(WZ\right)_{ij}=\mathrm{Re}\left(\frac{F_{W_i}}{Z_j}\right),\hfill \\ & & & & & \\ \hfill \left(WZ\right)_{ij}^{}& =& \mathrm{Re}\left(\frac{F_{W_i}}{Z_j^{}}\right),\hfill & \hfill F_{W_i}& =& \frac{s^{}}{W_i^{}}\frac{s^{}k_+tk_{}^{}+uk_+^{}4i\epsilon (q_+q_{}p_+k)}{2W_iW_i^{}},\hfill \end{array}$$ (18) using the notation $`\epsilon (p_1p_2p_3p_4)=\epsilon _{\mu \nu \rho \sigma }p_1^\mu p_2^\nu p_3^\rho p_4^\sigma `$, where $`\epsilon _{\mu \nu \rho \sigma }`$ is the completely antisymmetric Levi-Civita tensor and $`\epsilon _{0123}=1`$. In Eq. (17) we have assumed lepton universality with regards to the $`Z^{}\nu \overline{\nu }`$ couplings. Although it may not be immediately apparent, the contribution to the cross section from the state where the $`e^{}`$ and $`e^+`$ are both left-handed is equal to the contribution from the state where they are both right-handed and the sum is given by the term in Eq. (17) proportional to $`a_{W_i}a_{W_j}b_{W_i}b_{W_j}`$. A relation which was quite useful in simplifying $`||^2`$ is $$\frac{s^{}}{W_i^{}}\frac{s^{}k_+tk_{}^{}+uk_+^{}4i\epsilon (q_+q_{}p_+k)}{2W_iW_i^{}}=\frac{s^{}}{W_i}\frac{s^{}k_{}t^{}k_+^{}+u^{}k_{}^{}4i\epsilon (q_{}q_+p_{}k)}{2W_iW_i^{}}.$$ (19) In the SM limit, Eq. (17) agrees with the expression given in after correcting for the known missing factors of $`1/s^{}`$ in required on dimensional grounds. The calculation of $`d\sigma /dE_\gamma d\mathrm{cos}\theta _\gamma `$ may be performed analytically or numerically. We have followed both approaches and verified numerical agreement. Further checks were performed using the program CompHEP . ## IV Results Before discussing the discovery limits obtained in the various models, we present the total cross sections and the differential cross sections $`d\sigma /dE_\gamma `$ and $`d\sigma /d\mathrm{cos}\theta _\gamma `$. In doing so, all the essential features are illustrated. We take the SM inputs $`M_W=80.33`$ GeV, $`M_Z=91.187`$ GeV, $`\mathrm{sin}^2\theta _W=0.23124`$, $`\alpha =1/128`$, $`\mathrm{\Gamma }_Z=2.49`$ GeV . Since we work only to leading order in $`||^2`$, there is some arbitrariness in what to use for the above input, in particular $`\mathrm{sin}^2\theta _W`$. Kinematically, the maximum allowed value for $`E_\gamma `$ is $`\sqrt{s}/2`$. In addition, to take into account detector acceptance, $`E_\gamma `$ and $`\theta _\gamma `$ have been restricted to the ranges $$E_\gamma 10\mathrm{GeV},10^0\theta _\gamma 170^0.$$ (20) The cuts also serve to remove the singularities which arise when the emitted photon is soft or collinear with the beam. Further, we restrict the photon’s transverse momentum to $$p_T^\gamma >\frac{\sqrt{s}\mathrm{sin}\theta _\gamma \mathrm{sin}\theta _v}{\mathrm{sin}\theta _\gamma +\mathrm{sin}\theta _v},$$ (21) where $`\theta _v`$ is the minimum angle down to which the veto detectors may observe electrons or positrons. We take $`\theta _v=25`$ mrad. This cut has the effect of removing the largest background to our process, namely radiative Bhabha-scattering where the scattered $`e^+`$ and $`e^{}`$ go undetected down the beam pipe. This study was performed in leading order, but QED corrections to $`e^+e^{}\nu \overline{\nu }\gamma `$ must be taken into account in a precision analysis of real data. They have been known to $`O(\alpha )`$ for some time . See for a short review of and further references to higher order QED corrections, and for a description of a related MC generator. Since our aim is to determine the statistical power of the process in discovering $`W^{}`$’s, there is no need to include in this study the radiative corrections which will only marginally influence the number of events. Complete consistency at NLO, however, would require determination of the bremsstrahlung corrections to the generalized expression (17) and corresponding loop graphs. As well, we do not explicitly take into account any higher order backgrounds. A background, which cannot be suppressed, comes from the reaction $`e^+e^{}\nu \overline{\nu }\nu ^{}\overline{\nu }^{}\gamma `$. The authors of have provided the following cross sections of relevance here: $`\sigma (e^+e^{}\nu _e\overline{\nu }_e\nu _e\overline{\nu }_e\gamma )\sigma _{eeee}=6.65(2)\mathrm{fb},\sigma _{ee\mu \mu }=7.79(2)\mathrm{fb},\sigma _{\mu \mu \mu \mu }=0.690(2)\mathrm{fb}`$ and $`\sigma _{\mu \mu \tau \tau }=1.383(3)\mathrm{fb}`$. These results are for the same conditions as in Table 1 of but for $`\sqrt{s}=500\mathrm{GeV}`$. The cuts used in obtaining the above numbers differ from ours. Nonetheless, these cross sections give an idea of the magnitude of the background. Assuming lepton universality, the total cross section is $`25\mathrm{fb}`$ for the process $`e^+e^{}\nu \overline{\nu }\nu ^{}\overline{\nu }^{}\gamma `$. Imposing our $`p_T^\gamma `$ cut will suppress it even further. This background must be included in an “$`e^+e^{}\gamma `$ \+ nothing” analysis of real data. We expect that the cross sections of $`e^+e^{}\nu \overline{\nu }\nu ^{}\overline{\nu }^{}\gamma \gamma `$ and of $`e^+e^{}\nu \overline{\nu }\nu ^{}\overline{\nu }^{}\nu ^{\prime \prime }\overline{\nu }^{\prime \prime }\gamma `$ are so small that they need not be taken into account in the analysis. The errors generated from the subtraction of the above backgrounds form part of the systematic error. As the backgrounds themselves are much smaller than the signal, though comparable to the new physics effect, we expect that the error in the SM prediction of the backgrounds would be much smaller than the systematic errors arising from detector and beam uncertainties. We shall return to the issue of systematics in connection with their influence on the discovery limits presented in the next section. We have calculated three distinct total cross sections: unpolarized: $`\sigma `$, for left-handed $`e^{}`$: $`\sigma _L`$, and for right-handed $`e^{}`$: $`\sigma _R`$. Fig. 2 shows all three plotted versus $`\sqrt{s}`$, with $`\sigma _L`$ and $`\sigma _R`$ calculated using 100% beam polarization. Results are shown for the SM, LRM ($`\rho =\kappa =1`$), UUM ($`\mathrm{sin}\varphi =0.6`$), SSM($`W^{}`$), SSM($`W^{}+Z^{}`$) and KK model, with $`M_W^{}=750`$ GeV in each case. These mass and coupling parameter choices are rather arbitrary, made to illustrate general behaviour. It is worth noting at this point that in the UUM and SSM($`W^{}+Z^{}`$), the correction to the SM cross section changes sign as $`\sqrt{s}`$ is varied. This arises, for certain $`\sqrt{s}`$ and $`M_W^{}`$, due to a negative interference term between the SM and $`Z^{}/W^{}`$ diagrams in these models. It is clear from the presence of the peaks in Fig. 2 that we are also probing $`Z^{}`$’s, in those models which include them. (There is also a very sharp peak at lower $`\sqrt{s}`$, off the plot, due to the SM $`Z`$.) The $`Z^{}`$ peaks generally occur for $`\sqrt{s}`$ slightly above the $`Z^{}`$ mass since the photon carries away some of the energy. At very high energies, the SM $`Z`$ contribution is negligible. Further, by using a right-handed $`e^{}`$ beam, we can reduce the SM $`W`$ contribution (depending on the degree of polarization). Then we directly probe the $`W^{}`$ (and $`Z^{}`$) in the LRM, while in the SSM($`W^{}+Z^{}`$) and KK model, we probe only the $`Z^{}`$. The latter two models as well as the two remaining models all require some component of left-handed polarization to probe the $`W^{}`$. The above features are borne out in Fig. 2. In order to see which regions of $`E_\gamma `$ are most sensitive to the new physics, we plot for left- and right-handed electron beams respectively, in Figs. 3(a) and 4(a) $`d\sigma /dE_\gamma `$ versus $`E_\gamma `$ and in Figs. 3(b) and 4(b) the deviation from the SM result divided by the square root of the predicted cross section versus $`E_\gamma `$. We show results for $`\sqrt{s}=500`$ GeV with 100% $`e^{}`$ beam polarization in these figures. First, we note the shape of $`d\sigma /dE_\gamma `$ in Figs 3(a) and 4(a). For left-handed electrons, the bulk of the cross section comes from the low $`E_\gamma `$ region; the reduction at very low $`E_\gamma `$ is due to the $`p_T^\gamma `$ cut and the sharp peak at $`E_\gamma 240`$ GeV is due to the radiative return to the $`Z`$ pole. For 100% right polarized electrons, the cross section is rather flat in the low to moderate $`E_\gamma `$ region, then increases as a result of the $`Z`$ peak at high $`E_\gamma `$. On the other hand, since the right-handed cross section is two orders of magnitude smaller than the left-handed cross section away from the $`Z`$ peak, any realistic degree of polarization (i.e. 90%) will lead to a large contribution from $`\sigma _L`$ to the low $`E_\gamma `$ region. In general, there can also be a peak due to a $`Z^{}`$ for $`M_Z^{}<\sqrt{s}`$ which occurs at $$E_\gamma ^{\mathrm{peak}}=\frac{\sqrt{s}}{2}\left(1\frac{M_Z^{}^2}{s}\right)$$ (22) in analogy with the SM $`Z`$. Most important, however, is the relative statistical significance, shown in Figs. 3(b) and 4(b). In both the left- and right-handed cases, the low $`E_\gamma `$ region is the most sensitive to the new physics. There are two reasons for this. First, for left-handed electrons, the cross section is largest at low $`E_\gamma `$, as mentioned above. Second, the lower $`E_\gamma `$, the higher the mass probed in the $`Z^{}`$ propagator via Eq. (8). The relative effect is even larger when combining the $`\chi ^2`$’s from the different bins, since it is the squares of the plotted quantities which will enter. Overall, the KK model leads to the most statistically significant deviations, except for the 100% left polarized case where the SSM($`W^{}`$) exhibits the largest deviation. We can also see clearly how the sign of the deviation from the SM depends on the beam polarization. For the KK model and SSM($`W^{}+Z^{}`$), we observe a negative deviation with right-handed polarization, implying a negative $`Z^{}`$ contribution, versus a positive overall contribution coming from the left-handed channel. Clearly, interference effects will make probing $`W^{}`$’s nontrivial. We shall return to this point in the next section. In Figs. 5 and 6 we plot the analogous quantities relevant to $`d\sigma /d\mathrm{cos}\theta _\gamma `$, versus $`\mathrm{cos}\theta _\gamma `$. We note that both $`d\sigma /d\mathrm{cos}\theta _\gamma `$ and the relative statistical significance are peaked in the forward and backward directions and both are very nearly symmetric in $`\mathrm{cos}\theta _\gamma `$. The latter implies that the forward-backward asymmetry will be small and, therefore, the deviation from the SM forward-backward asymmetry will also be small, at least in absolute magnitude. We therefore do not expect the forward-backward asymmetry to serve as a useful probe of the new physics, which is confirmed by explicit calculation. An important observation is that our $`p_T^\gamma `$ cut, while eliminating a large background, has also eliminated much of our signal (both from the small angle and soft events) which was appreciably stronger prior to the cut. A more detailed study, including a detector simulation, would be required to determine whether the background could be accurately subtracted with a looser $`p_T^\gamma `$ cut. ### A Discovery Limits for $`W^{}`$’s The best discovery limits were in general obtained using the observable $`d\sigma /dE_\gamma `$, combined with beam polarization, while $`d\sigma /d\mathrm{cos}\theta _\gamma `$ was less sensitive. Comparable or equal limits were obtained using the total cross section, with an additional cut on the energy to eliminate the $`Z`$ pole radiative return events: $$E_\gamma ^{\mathrm{max}}=\frac{\sqrt{s}}{2}\left(1\frac{M_Z^2}{s}\right)6\mathrm{\Gamma }_Z.$$ (23) As can be seen from Figs. 3(b) and 4(b), the $`Z`$ pole region is quite insensitive to new physics. In the cases that $`d\sigma /dE_\gamma `$ provided a better limit than the total cross section, the improvement was of order 50 GeV. However, the $`\chi ^2`$ obtained using the total cross section is a somewhat less stable function of $`M_W^{}`$ as the sign of the deviation from the SM cross section may change with $`M_W^{}`$ leading to isolated regions of insensitivity at low $`M_W^{}`$. Also, when systematic errors are included, the limits obtained using $`d\sigma /dE_\gamma `$ are affected much less than those obtained using the total cross section. Substantially weaker limits were obtained using the left-right asymmetry, $$A_{LR}=\frac{\sigma _L\sigma _R}{\sigma _L+\sigma _R},$$ (24) even when including systematic errors only one half those used in the $`d\sigma /dE_\gamma `$ calculation (since one expects some cancellation of errors between the numerator and denominator in $`A_{LR}`$). As expected from the discussion of the previous section, the forward-backward asymmetry, $`A_{FB}`$, was quite insensitive to the new physics. In light of the above, we restrict the remaining discussion to limits obtained using $`d\sigma /dE_\gamma `$ as an observable. In obtaining the $`\chi ^2`$ for $`d\sigma /dE_\gamma `$, we used 10 equal sized energy bins in the range $`E_\gamma ^{\mathrm{min}}<E_\gamma <E_\gamma ^{\mathrm{max}}`$, where $`E_\gamma ^{\mathrm{min}}`$ follows from the $`p_T^\gamma `$ cut Eq. (21): $$E_\gamma ^{\mathrm{min}}=\frac{\sqrt{s}\mathrm{sin}\theta _v}{1+\mathrm{sin}\theta _v},$$ (25) which supersedes the acceptance cut of Eq. (20). We have $$\chi ^2=\underset{\mathrm{bins}}{}\left(\frac{d\sigma /dE_\gamma d\sigma /dE_{\gamma ,\mathrm{SM}}}{\delta d\sigma /dE_\gamma }\right)^2,$$ (26) where $`\delta d\sigma /dE_\gamma `$ is the error on the measurement and analogous formulae hold for other observables. One sided 95% confidence level discovery limits are obtained by requiring $`\chi ^22.69`$ for discovery. Systematic errors, when included, were added in quadrature with the statistical errors. In determining the limits for the case of polarized electron beams, we show results for the polarization state which in general has the largest sensitivity (deviation from the SM) for a given model; a right-handed $`e^{}`$ beam for the LRM and a left-handed beam for all other models. We used one half the unpolarized luminosity for the polarized case, assuming equal running time in each polarization state. The discovery limits for all five models are listed in Table I, for $`\sqrt{s}=0.5`$, 1.0 and 1.5 TeV, using the same input parameters as for the cross sections presented in the previous section. We show limits for both an unpolarized $`e^{}`$ beam and for a 90% polarized one. For each center-of-mass energy, two luminosity scenarios are considered and we present limits obtained with and without systematic errors. Our prescription is to include a 2% systematic error per bin. This number is quite arbitrary but seems reasonable, if not conservative, considering the clean final state. In addition to detector systematics, which we expect will dominate, there are uncertainties associated with the beam luminosity and energy, which will be spread over a range. The systematic errors associated with the background subtraction should be much smaller than 2% as should be the errors in the calculation of the QED corrections. The 2% number should not be taken too seriously therefore, except to highlight the fact that a precision measurement is required to take full advantage of the large event rate. Certain features are common to all models. With no systematic error included, we observe quite an improvement in the limits with increased luminosity. The only exception is the UUM at $`\sqrt{s}`$ of 1.5 TeV, where the improvement is minimal. The reason is that the $`\chi ^2`$ decreases very rapidly as $`M_W^{}`$ is increased in the vicinity of the limit, hence increasing the luminosity by a factor of 2.5 does little. The unusual $`\sqrt{s}`$ dependence can be attributed to the interference effect noted in the previous section, which results in, for example, for the UUM with $`\mathrm{sin}\varphi =0.6`$ and an integrated luminosity of 500 fb<sup>-1</sup>, a lower discovery limit at $`\sqrt{s}=1.5`$ TeV than at 0.5 and 1 TeV. We will return to this peculiar behaviour later in the section. When 2% systematic errors are included, the high luminosity scenario yields little improvement in the limits in any of the models, since the systematic error now dominates the statistical. Perhaps surprising at first is the observation that 90% beam polarization does not improve the limits very much. This follows from taking into account the reduced luminosity and the fact that the left-handed component tends to dominate the unpolarized cross section by a considerable amount. On the other hand, we observed that if the polarization is pushed beyond 90%, then the right-polarized limits can increase significantly in those models in which the beyond-SM bosons have a non-zero right-handed coupling: the LRM, KK model and SSM($`W^{}+Z^{}`$). In the latter two models, it is however, the $`Z^{}`$ which is being probed. The higher degree of polarization is required to eliminate the contamination from the much larger left-handed component. Thus, the primary advantage of beam polarization is to distinguish between models and measure the new couplings, as will be investigated in the next section. Fig. 7 presents the $`W^{}`$ mass discovery limits obtainable in the LRM with an unpolarized beam, plotted versus $`\kappa `$ for $`\rho =1`$ and $`\sqrt{s}=`$ $`0.5`$, 1.0, 1.5 and 2 TeV using a luminosity of 50 fb<sup>-1</sup> for $`\sqrt{s}=0.5`$ TeV and 200 fb<sup>-1</sup> for the higher energies. Only statistical errors are included. Depending on $`\sqrt{s}`$ and $`\kappa `$, the limits range from 0.8 to 2.8 TeV. We expect greater deviations from the SM, and hence larger limits, as $`\kappa `$ is increased since this increases the $`W^{}`$ coupling strength, as can be seen from Eq. (3). The predicted dependence on $`\kappa `$ is generally observed, except at low $`\kappa `$ where we notice a moderate increase in the limits, even though the $`W^{}`$ couplings have weakened. We attribute this effect to the $`Z^{}`$, whose couplings are enhanced (but its mass increased) in the low $`\kappa `$ region. This was indicated by an appreciable improvement in the limits for low $`\kappa `$ and $`\rho =1`$ versus those obtained using $`\rho =2`$ and consequently a heavier $`Z^{}`$, via Eq. (1). Fig. 8 demonstrates the improvement in bounds in the moderate to large $`\kappa `$ region obtained when a 90% or 100% polarized right-handed $`e^{}`$ beam is used. The beam polarization picks out the LRM $`W^{}`$ and suppresses the SM $`W`$. Fig. 8(a) shows that for $`\kappa >1`$, 90% beam polarization improves the limits. Further increasing the polarization leads to substantial improvements, even at lower $`\kappa `$, as demonstrated in Fig. 8(b). The dependence of the limits in the UUM on $`\mathrm{sin}\varphi `$ is shown in Fig. 9, for $`\sqrt{s}=`$ $`0.5`$, 1.0, 1.5 and 2.0 TeV, under the same running conditions as Fig. 7. Only the unpolarized case is considered as beam polarization was not beneficial. Again, only statistical errors are included. At each $`\sqrt{s}`$, we note that the contour defining the exclusion region as a function of $`\mathrm{sin}\varphi `$ is a complicated curve. The consequence is that for $`\sqrt{s}=1`$ TeV, we obtain better limits over a range of $`\mathrm{sin}\varphi `$ than we do for $`\sqrt{s}=1.5`$ and even $`\sqrt{s}=2`$ TeV. Essentially, this is due to the complicated interference with the SM diagrams. In general, as $`\mathrm{sin}\varphi `$ increases, the UUM couplings also increase, as can be seen from Eq. (4), so that higher mass scales are probed. So, referring to Fig. 2, the peak in the cross section (due to the $`Z^{}`$) at the scale being probed shifts to the right. But the sign of the deviation from the SM changes with $`\sqrt{s}`$ for fixed $`M_W^{}`$ (or vice-versa) such that the UUM cross section dips below the SM over some region to the left of the peak, then goes back above it for small $`\sqrt{s}`$ (or large $`M_W^{}=M_Z^{}`$ for fixed $`\sqrt{s}`$). Hence, there is a small step in the limits near $`M_W^{}=\sqrt{s}`$, corresponding to passing the rightmost crossing with the SM and another structure in the contour at some higher $`M_W^{}`$ such that the leftmost crossing is situated near $`\sqrt{s}`$. One sees this explicitly by plotting $`\chi ^2`$ versus $`M_W^{}`$ for fixed $`\sqrt{s}`$ and $`\mathrm{sin}\varphi `$ and observing a dip in the $`\chi ^2`$ at relatively low $`M_W^{}`$. Had we used $`\sigma `$ as an observable, the dip would be much more pronounced since $`\sigma \sigma _{\mathrm{SM}}`$ passes through zero, but $`d\sigma /dE_\gamma d\sigma /dE_{\gamma ,\mathrm{SM}}`$ may differ in sign between bins, leading to a nonzero $`\chi ^2`$ at the crossing points. Once $`\mathrm{sin}\varphi `$ is large enough that we are probing the region to the left of the leftmost crossing, the limits shoot up in an impressive fashion as the dip in $`\chi ^2`$ never goes back down to 2.69. The shape of the plot is luminosity dependent since, as pointed out earlier in this section, the degree to which increased luminosity improves the limits depends on the rate at which the $`\chi ^2`$ decreases with increasing $`M_W^{}`$ in the vicinity of the limit. That, in turn, varies with $`\sqrt{s}`$ for fixed $`\mathrm{sin}\varphi `$ and with $`\mathrm{sin}\varphi `$ for fixed $`\sqrt{s}`$. ### B Constraints on Couplings In this section, we consider constraints which can be put on the couplings of extra gauge bosons by the process $`e^+e^{}\nu \overline{\nu }\gamma `$. These constraints are significant only in the case where the mass of the corresponding extra gauge boson is considerably lower than its search limit in this process. In most models, the process $`e^+e^{}f\overline{f}`$ and/or searches at the LHC are more sensitive to a $`Z^{}`$ or $`W^{}`$ (LHC) than the process $`e^+e^{}\nu \overline{\nu }\gamma `$. We assume here that a signal for an extra gauge boson has been detected by another experiment. Given such a signal, we derive constraints (at 95% C.L.) on the couplings of extra gauge bosons. We present the constraints in terms of couplings normalized as follows relative to Eqs. (11) and (12). $$\begin{array}{cccc}\hfill L_f(Z)=& \frac{g}{4c_W}a_{Z_i}^f\hfill & \hfill R_f(Z)=& \frac{g}{4c_W}b_{Z_i}^f\hfill \\ \hfill L_f(W)=& \frac{g}{2\sqrt{2}}a_{W_i}\hfill & \hfill R_f(W)=& \frac{g}{2\sqrt{2}}b_{W_i}.\hfill \end{array}$$ (27) The constraints correspond to $$\chi ^2=\underset{i}{}\left(\frac{O_i(SM)O_i(SM+Z^{}+W^{})}{\delta O_i}\right)^2=5.99,$$ (28) where $`O_i(SM)`$ is the prediction for the observable $`O_i`$ in the SM, $`O_i(SM+Z^{}+W^{})`$ is the prediction of the extension of the SM and $`\delta O_i`$ is the expected experimental error. The index $`i`$ corresponds to different observables such as $`\sigma `$ and $`A_{LR}`$. Our assumptions concerning beam polarization are as follows. For single beam ($`e^{}`$) polarization, we assume, as in the previous section, equal running in left and right polarization states. For double beam polarization, we assume equal running in the $`LR`$ and $`RL`$ states, but no running in the $`LL`$ and $`RR`$ states. Thus, $$A_{LR}=\frac{\sigma _{LR}\sigma _{RL}}{\sigma _{LR}+\sigma _{RL}},e^{}\text{ and }e^+\text{ polarized},$$ (29) where the first subscript of $`\sigma `$ refers to the $`e^{}`$ helicity. Note that for 100% polarized $`e^{}`$ and $`e^+`$, $`\sigma _{LL}=\sigma _{RR}=0`$ in all the models we consider. This remains approximately valid as the couplings deviate from their model-defined values. In Figs. 10 and 11, we present $`Z^{}\nu \overline{\nu }`$ coupling constraints assuming there is no signal for a $`W^{}`$. This is the case when the SM is extended by $`U(1)`$ factors only. It can also happen in models where the $`W^{}`$ has purely right-handed couplings and the right-handed neutrino is heavy. Then, the process $`e^+e^{}\nu \overline{\nu }\gamma `$ would be one of the best for constraining the couplings of the $`Z^{}`$ to SM neutrinos below the $`Z^{}`$ resonance. If there is also a signal for a $`W^{}`$, a similar analysis could be performed including the $`W^{}`$ parameters, as measured in other experiments. The resulting bounds would be larger than those shown in the two figures. However, the main points of the discussion would remain unchanged. Fig. 10 illustrates the resulting constraints on a 1.5 TeV $`Z^{}`$ at a 500 GeV collider for different observables and experimental parameters, including luminosity and beam polarization. We see that we can get some interesting constraints even though the $`Z^{}`$ is considerably heavier than the centre-of-mass energy. The region which cannot be resolved by the observables is between the two corresponding lines and contains the couplings of the SM. Hence the star in this figure corresponds to the SSM($`Z^{}`$). For the cases where only one bounding line is shown, the second line is outside the figure. $`R_\nu (Z^{})`$ and $`L_\nu (Z^{})`$ are mainly constrained by the interference of the $`Z^{}`$ exchange with the SM. The strongest constraint is on the $`Z^{}`$ coupling to left-handed neutrinos. This makes the constraints especially simple. First we consider an integrated luminosity of 500 fb<sup>-1</sup>. The total unpolarized cross section gives the strongest constraint. The constraints from energy and angular distributions (with 10 equal size bins) were also considered but they give no improvement. The constraint from $`A_{LR}`$ is shown for two polarization cases: 90% electron beam polarization and the case of a collider with a $`P^{}=90\%`$ polarized electron beam and a $`P^+=60\%`$ polarized positron beam. Even for the latter case, the constraint from $`A_{LR}`$ is worse than that from the total cross section. We mention here for completeness that two polarized beams give not only a high effective polarization but also a small effective polarization error . The constraint obtained with an integrated luminosity of $`L_{\mathrm{int}}=50`$ fb<sup>-1</sup> is also shown in Fig. 10, to contrast with the high luminosity case. We see that for $`L_{\mathrm{int}}=500`$ fb<sup>-1</sup> a systematic error of 1% relaxes the constraints considerably and dilutes the advantage of high luminosity. Thus, both small systematic errors and a high luminosity collider are highly desired for the proposed measurement. Fig. 11 shows the possible constraints on $`R_\nu (Z^{})`$ and $`L_\nu (Z^{})`$ from $`\sigma `$ and $`A_{LR}`$, including systematic errors, for two representative $`Z^{}`$ masses, 0.75 TeV and 1 TeV. The constraints become much stronger as the $`Z^{}`$ mass is decreased. So far, we assumed that the $`Z^{}e^+e^{}`$ couplings, $`R_e(Z^{})`$ and $`L_e(Z^{})`$, are precisely known. However, they must be measured (with errors) by another experiment. Fig. 4(b) of illustrates such a measurement for a collider with a luminosity of 20 fb<sup>-1</sup>. To estimate their influence on the $`R_\nu (Z^{})`$, $`L_\nu (Z^{})`$ constraint, we make use of the errors on $`R_e(Z^{})`$ and $`L_e(Z^{})`$ given in . Our input for the errors of the $`Z^{}e^+e^{}`$ couplings for $`M_Z^{}=1.0`$ TeV and $`0.75`$ TeV are obtained from those for $`1.5`$ TeV by the scaling relation (2.63) in . We see that the uncertain knowledge of the $`Z^{}e^+e^{}`$ couplings leads to only slightly weaker constraints on $`R_\nu (Z^{})`$ and $`L_\nu (Z^{})`$. However, Fig. 11 shows that this effect is only important for a relatively heavy $`Z^{}`$ and for $`R_\nu (Z^{})`$ (even at lower $`Z^{}`$ masses) for which the constraints are already weak. Finally, we mention that there is no sign ambiguity in the measurement of $`R_\nu (Z^{})`$ and $`L_\nu (Z^{})`$ if the signs of the $`Z^{}e^+e^{}`$ couplings are known. It was noted that the $`Z^{}e^+e^{}`$ couplings have a two-fold sign ambiguity if measured in the process $`e^+e^{}e^+e^{}`$ alone. If this ambiguity exists, it induces a related sign ambiguity for $`R_\nu (Z^{})`$ and $`L_\nu (Z^{})`$. If the sign ambiguity in the $`Z^{}e^+e^{}`$ couplings is resolved (i.e. by measurements obtained from the process $`e^+e^{}W^+W^{}`$ below the $`Z^{}`$ resonance or by measurements at the $`Z^{}`$ resonance) it also disappears in our constraints on $`R_\nu (Z^{})`$ and $`L_\nu (Z^{})`$. In Figs. 12 to 15, we shall assume that there is no signal from a $`Z^{}`$ but that a signal from a $`W^{}`$ has been observed. This could happen in models where the $`W^{}`$ is considerably lighter than the $`Z^{}`$. We recognize that this particular scenario is unlikely in the context of the models we consider. For instance, in the UUM, the $`W^{}`$ and $`Z^{}`$ masses are approximately equal and there would most likely be a signal observed for the $`Z^{}`$ in addition to the $`W^{}`$. The situation is similar in the LRM, where the relationship between the $`W^{}`$ and $`Z^{}`$ masses is given in Eq. (1). Thus, it should be understood that our results for the case of a $`W^{}`$ only represent an estimate of the reach of this process in constraining $`W^{}`$ couplings, rather than precision limits in the context of a full understanding of the physics realized in nature. We use this simple scenario in order to indicate sensitivity to various parameters, such as the observables used and the luminosity. Alternatively, a known $`Z^{}`$ could be included in the following analysis. Again, the experimental errors on the measured $`Z^{}`$ parameters would enlarge the errors of the $`W^{}`$ measurements but not change the main conclusions. We will see that the process $`e^+e^{}\nu \overline{\nu }\gamma `$ can give model independent constraints on the quantities $`L_l(W^{})`$ and $`R_l(W^{})`$ for $`W^{}`$ masses considerably larger than the center-of-mass energy. We only probe $`l=e`$ directly, but we are assuming lepton universality throughout. Fig. 12 is similar to Fig. 10, but it shows the constraints on the $`W^{}`$ couplings. In this figure, for illustration, we assume there exists a $`W^{}`$ with SM couplings but with a mass of 1.5 TeV and that the right-handed neutrino is light enough to be produced. We find that the left- and, to some extent, the right-handed $`W^{}`$ coupling can be constrained. The figure illustrates the use of different combinations of $`\sigma `$ and $`A_{LR}`$, and of different beam polarizations. The unpolarized cross section mainly constrains the left-handed $`W^{}`$ coupling because left-handed electrons give its dominant contribution. The constraints from energy and angular distributions give almost no improvement for the model considered here. The constraint from $`A_{LR}`$ is complementary to that from $`\sigma `$. It is shown for the two cases of 90% electron beam polarization and for 90% electron beam polarization with 60% positron polarization. We see that $`\sigma `$ and $`A_{LR}`$ together give the best constraints on the couplings. The constraints on the $`W^{}`$ couplings have a two-fold sign ambiguity; nothing is changed by a simultaneous change of the sign of $`L_l(W^{})`$ and $`R_l(W^{})`$. The reason for this ambiguity lies in the squared amplitude, Eq. (17), where these couplings always enter as squares or as a product of left and right $`W^{}`$ couplings. In the case where we have only a weak $`W^{}`$ signal, the two regions allowed by this ambiguity overlap into one large region. In Fig. 13, we show constraints on the $`W^{}`$ couplings from $`\sigma `$ and $`A_{LR}`$ combined. In this figure, we illustrate the use of different luminosities and the inclusion of a systematic error. We have the same two well separated regions for the case of high luminosity and no systematic error as in Fig. 12. These two regions become larger for low luminosity and no systematic error. We are left with one large region after the inclusion of a systematic error of 2% for $`\sigma `$ and 1% for $`A_{LR}`$. As in the case of extra neutral gauge bosons, small systematic errors and high luminosity are necessary for a coupling measurement. In Fig. 14, we show how the constraints on the $`W^{}`$ couplings vary for different $`W^{}`$ masses. The constraint for $`M_W^{}=1.5`$ TeV is identical to that from Fig. 13. We see that the constraint on the $`W^{}`$ couplings improves dramatically for lower $`W^{}`$ masses. Fig. 15 illustrates the possibility of discrimination between different models. We see that a $`W^{}`$ with SM couplings ($`W_L^{}`$) can be separated from the SM. A $`W^{}`$ with pure right-handed couplings ($`W_R^{}`$) with a strength of the left-handed coupling of the SM $`W`$ cannot be distinguished from the SM case. Looking at the squared amplitude, Eq. (17), we see that the constraints shown in Figs. 12 to 15 are, to a good approximation, valid for the combinations $`L_l(W^{})/M_W^{}`$ and $`R_l(W^{})/M_W^{}`$, and not for the couplings and the mass separately. We have fixed the $`W^{}`$ mass here for illustrational purposes. If a $`W^{}`$ is found with a mass different from our assumptions, the constraint on its couplings can be found by the appropriate scaling of our results. So far, we considered model independent bounds on the couplings of a single extra gauge boson while neglecting the existence of other extra gauge bosons. However, typically, extra neutral and charged gauge bosons simultaneously influence the observables. We consider this situation for the LRM and the UUM. In Fig. 16, we consider the Left-Right symmetric model. For $`M_W^{}=0.75`$ TeV, Eq. (1) gives $`M_Z^{}=0.90(1.27)`$ TeV for $`\kappa =1`$ and $`\rho =1(2)`$. We show the constraints on the couplings of the $`W^{}`$ for $`\rho =1`$ obtained by two different fitting strategies. First, we ignore the $`Z^{}`$ completely, and second, we take the $`Z^{}`$ into account assuming exact knowledge of its couplings. We see that the two curves are quite close. The reason is that our process is not very sensitive to such a $`Z^{}`$. These two curves are very similar to those for the $`W_R`$ and the SM in Fig. 15 because we are not very sensitive to a right-handed $`W^{}`$. The case of $`\rho =2`$ predicts a heavier $`Z^{}`$, which produces constraints differing even less from each other than those for $`\rho =1`$, so we do not show them. To demonstrate how the constraints change for a larger signal, we repeated the same procedure with $`M_W^{}=550`$ GeV. This number (and the mass of the associated $`Z^{}`$) are at the edge of the present exclusion limit . Although the constraints improve a bit, they are still not very impressive. Fig. 17 is similar to Fig. 16 but here we consider the Un-Unified model. We examine the cases $`M_W^{}=M_Z^{}=0.75`$ TeV and $`M_W^{}=M_Z^{}=0.55`$ TeV. We show the constraints on the couplings of the $`W^{}`$ obtained using the same two fitting strategies described for Fig. 16. Even for masses of $`0.75`$ TeV, the two curves are better separated than in LRM. For masses of 0.55 TeV, the wrong fitting strategy gives a region which is outside the true $`W^{}`$ coupling. This shows that such a light $`Z^{}`$ cannot be ignored in the fitting procedure. The process $`e^+e^{}f\overline{f}`$ and searches in hadron collisions are more sensitive to $`Z^{}`$ discovery than $`e^+e^{}\nu \overline{\nu }\gamma `$. A $`Z^{}`$ signal will always be detected in the cases where the $`Z^{}`$ contribution is relevant for a $`W^{}`$ constraint from $`e^+e^{}\nu \overline{\nu }\gamma `$. This information from other experiments will be required for a reliable $`W^{}`$ constraint from $`e^+e^{}\nu \overline{\nu }\gamma `$. ## V Conclusions In this paper, we studied the sensitivity of the process $`e^+e^{}\nu \overline{\nu }\gamma `$ to extra gauge bosons. We used this process to find discovery limits and to see how well one could measure the couplings of extra gauge bosons that are expected in extensions of the standard model. For the discovery limits we focused on $`W^{}`$’s since one can put better limits on $`Z^{}`$’s from other processes, such as $`e^+e^{}f\overline{f}`$, while, on the other hand, no similar limits exist on $`W^{}`$’s. The highest reach was obtained by binning the $`d\sigma /dE_\gamma `$ distribution although comparable results were obtained using the total cross section after the $`Z`$ radiative return was eliminated. The discovery reach is typically in the 1-6 TeV range depending on the specific model, the center of mass energy, and the assumed integrated luminosity. These results are substantially degraded if one includes systematic errors. For the $`W_R`$ boson, for which LHC discovery limits are available, the discovery limits are, for $`g_R=g_L`$, $`M_W^{}`$= 1.2, 1.6, and 1.9 TeV for $`\sqrt{s}=`$ 500, 1000, and 1500 GeV respectively assuming $`L_{\mathrm{int}}=500`$ fb<sup>-1</sup> relative to a reach of $`5.9`$ TeV at the LHC. Although the discovery reach for $`W^{}`$’s of this process is not competitive with the reach of the LHC, precision measurements can give information on extra gauge boson couplings which complements the LHC. In particular, if the LHC were to discover a $`Z^{}`$ or $`W^{}`$ the process $`e^+e^{}\nu \overline{\nu }\gamma `$ could constrain $`Z^{}`$ and $`W^{}`$ couplings. For a $`Z^{}`$, this would be the best measurement of the $`Z^{}\nu \overline{\nu }`$ couplings. For $`W^{}`$ couplings, reliable measurements would require information from, for example, $`e^+e^{}f\overline{f}`$ and searches in hadron collisions which would always detect a $`Z^{}`$ signal in the cases where its contribution is relevant for a $`W^{}`$ constraint by $`e^+e^{}\nu \overline{\nu }\gamma `$. Finally, we emphasize that to make measurements of the extra gauge boson couplings, high luminosity will be needed and it will be very important to reduce the systematic uncertainties as much as possible. ###### Acknowledgements. This research was supported in part by the Natural Sciences and Engineering Research Council of Canada. S.G., P.K., and B.K. thank Dean Karlen and A.L. thanks Graham Wilson for useful discussions. ## appendix Here we give explicit parametrizations of the momenta defined in the frame where $`q_{},q_+`$ are back-to-back and $`p_+`$ defines the $`\widehat{z}`$ axis, suitable for use with the phase-space (9): $`p_+`$ $`=`$ $`(\omega _+;0,0,\omega _+)`$ (30) $`p_{}`$ $`=`$ $`(\omega _{};\omega _k\mathrm{sin}\psi ,0,\omega _k\mathrm{cos}\psi \omega _+)`$ (31) $`k`$ $`=`$ $`(\omega _k;\omega _k\mathrm{sin}\psi ,0,\omega _k\mathrm{cos}\psi )`$ (32) $`q_+`$ $`=`$ $`(\omega _+^{};\omega _+^{}\mathrm{sin}\theta \mathrm{cos}\phi ,\omega _+^{}\mathrm{sin}\theta \mathrm{sin}\phi ,\omega _+^{}\mathrm{cos}\theta )`$ (33) $`q_{}`$ $`=`$ $`(\omega _+^{};\omega _+^{}\mathrm{sin}\theta \mathrm{cos}\phi ,\omega _+^{}\mathrm{sin}\theta \mathrm{sin}\phi ,\omega _+^{}\mathrm{cos}\theta ),`$ (34) where $`\omega _{}`$ $`=`$ $`{\displaystyle \frac{sk_{}}{2\sqrt{s^{}}}},\omega _+={\displaystyle \frac{sk_+}{2\sqrt{s^{}}}},\omega _+^{}={\displaystyle \frac{\sqrt{s^{}}}{2}},`$ (35) $`\omega _k`$ $`=`$ $`{\displaystyle \frac{ss^{}}{2\sqrt{s^{}}}},\mathrm{cos}\psi ={\displaystyle \frac{sk_{}s^{}k_+}{(sk_+)(ss^{})}}.`$ (36) It is arbitrary whether $`\mathrm{sin}\psi `$ is taken as positive or negative as long as one is consistent.
warning/0001/astro-ph0001142.html
ar5iv
text
# Gas Rich Dwarf Spheroidals ## 1 Introduction Recently, Blitz et al. (1999) have shown that the HVCs, clouds of atomic hydrogen inconsistent with near-circular rotation about the Galactic Center, are well explained if they are members of the Local Group. A simple dynamical model can replicate both the observed distribution on the sky as well as the observed kinematics of the ensemble of the HVCs. In their paper, Blitz et al. (1999) made several predictions; among them are that HVCs have H$`\alpha `$ surface brightnesses less than those measured in the Magellanic Stream, metallicities of 0.1 solar or less, and internal pressures P/k $``$ 1-10 K $`\mathrm{cm}^3`$; the last of these is inferred from the self-gravity of HVCs. All of these predictions have been subsequently confirmed (Weiner et al. 2000; Wakker et al. 1999; Sembach et al. 1999). Only one set of observations now seems to present problems for the model: three deep, blind extragalactic $`\mathrm{H}\mathrm{I}`$ surveys, two made at Arecibo (Zwaan et al. 1997; Spitzak & Schneider 1998), and one at Parkes (Banks et al. 1999). If the HVCs are extragalactic and part of the Local Group, as Blitz et al. (1999) argue, then extragalactic analogues should be observable; many examples have indeed been reported in the literature (Blitz et al. 1999 and references therein). The blind $`\mathrm{H}\mathrm{I}`$ surveys have also found many uncatalogued $`\mathrm{H}\mathrm{I}`$ clouds, but on closer inspection, all but one of the detections are found to harbor galaxies, albeit often of very low surface brightness. Because stars have not been detected in HVCs, they have been assumed to be starless systems. No such systems have been found by either Zwaan et al. (1997) or Banks et al. (1999); and only one potentially starless system has been found in the Spitzak & Schneider (1998) survey. The sensitivity of both surveys appears to be high enough and the velocity coverage large enough, that at least a few HVC analogues should have been detected in these surveys. On the other hand, it is unclear whether the lowest surface brightness galaxies found in the optical follow-up to the $`\mathrm{H}\mathrm{I}`$ detections would have been detected if those systems were located in the Local Group. These galaxies, which compose a significant fraction of the Zwaan et al. (1997) and Spitzak & Schneider (1998) $`\mathrm{H}\mathrm{I}`$ identifications, appear morphologically similar to the Local Group dwarf spheroidal galaxies, which have been observed to be generally gas-free (Mateo 1998). At Local Group distances such low surface brightness systems would be extended and relatively difficult to identify in existing surveys. However, with concerted searching, a number of dSph galaxies have recently been discovered in the Local Group (Whiting et al. 1997; Armandroff, Davies, & Jacoby 1998; Karachentsev & Karachentseva 1999; Gallart et al. 1999), suggesting that there might be similar, or even lower surface brightness galaxies associated with HVCs. A search for low surface brightness galaxies in HVCs is currently underway, but as a first step, we decided to examine the newly discovered Local Group dwarfs for 21-cm emission using the Leiden-Dwingeloo $`\mathrm{H}\mathrm{I}`$ survey (LDS - Hartmann & Burton 1997) to see if any are associated with HVCs. After the detection of $`\mathrm{H}\mathrm{I}`$ toward And V, and its subsequent identification as HVC 368 in the compilation of Wakker & van Woerden (1991), we decided to reexamine the $`\mathrm{H}\mathrm{I}`$ content of all of the dSph galaxies in the Local Group; the results are presented in §3. We have found $`\mathrm{H}\mathrm{I}`$ toward four galaxies in which it had not previously been detected, and found $`\mathrm{H}\mathrm{I}`$ more extended than previously thought in two others. The $`\mathrm{H}\mathrm{I}`$ found toward two galaxies are catalogued HVCs. In §4 we examine the implications of the $`\mathrm{H}\mathrm{I}`$ observations and infer the existence of a hot gaseous corona around the Milky Way and M31 with a mean density of $`2.5\times 10^5`$ $`\mathrm{cm}^3`$. Rather than being gas-poor, the dSph galaxies are often gas-rich, but with rather extended $`\mathrm{H}\mathrm{I}`$ envelopes. ## 2 Analysis The sensitivity of the LDS is about 70 mK in a 1 $`\mathrm{km}\mathrm{s}^1`$ velocity channel; its angular resolution is 36′. The survey covers the entire northern sky down to a declination of -30 at a sampling interval of 30′. At a distance of 100 kpc, the beam is 1.05 kpc. Its effective velocity coverage of -450 $`\mathrm{V}_{\mathrm{LSR}}`$ +400 $`\mathrm{km}\mathrm{s}^1`$ is sufficient to detect all Local Group emission down to a 5$`\sigma `$ column density N(H)/$`(\mathrm{\Delta }V)^{1/2}=6.4\times 10^{17}`$ $`\mathrm{cm}^2`$ ($`\mathrm{km}\mathrm{s}^1`$)<sup>-1/2</sup> averaged over the beam ($`\mathrm{\Delta }V`$ is the full width at half maximum of the $`\mathrm{H}\mathrm{I}`$ emission). The survey also has the virtue of having flat baselines which makes it possible to detect very low level emission at velocities away from the normal Galactic emission. We initially examined a five point cross centered at the optical position of all of the galaxies catalogued as dSph or dSph/dIrr in the compilation of Mateo (1998), but excluded those also classified as dE or E systems because of their much higher surface brightness. We also examined the galaxies subsequently identified as possible Local Group dSph galaxies by Karachentsev & Karachentseva (1999) and Gallart et al. (1999) which could be found in the LDS. If a galaxy sits between grid points of the survey, we examined a somewhat larger area. A number of the galaxies were found to have 21-cm emission confined to a small area coincident with or very close to the dSph at velocities outside the range normally associated with Galactic emission. One can estimate the probability of a chance coincidence of a dSph with a cloud along the line of sight from the surface filling fraction of small HVCs ($``$ 1 deg<sup>2</sup>) in the Wakker & van Woerden (1991) compilation. The HVCs comprise all of the $`\mathrm{H}\mathrm{I}`$ emission not associated with galaxies outside the range of normal Galactic emission and have the same range of radial velocities as the dSphs. The total area on the sky of small HVCs is less than 300 deg<sup>2</sup>, though this number is somewhat uncertain at the 50% level because of the sparse sampling of the Wakker & van Woerden (1991) catalogue. It is certainly an upper limit, though, based on the higher resolution mapping of about 20% of the smaller HVCs by Blitz et al. (1999) and by Braun & Burton (1999). The HVCs are spread all over the sky with some concentration in the general directions of the barycenter and antibarycenter of the Local Group, thus the probability of a chance coincidence with a cloud that subtends an angle of less than 1 deg<sup>2</sup> is $`0.01`$. Equivalently, a positional coincidence with such a cloud is significant at the $`2.5\sigma `$ level. We therefore consider the likelihood of a chance coincidence of a dSph with a small $`\mathrm{H}\mathrm{I}`$ cloud to be sufficiently low that it is indicative of a real physical association, even if no velocity is available for the galaxy. Indeed, in a sample of $``$20 objects such as considered here, the probability of a chance positional coincidence of any dSph with an $`\mathrm{H}\mathrm{I}`$ cloud is $``$0.2, and the number of detections is about 50 times higher. In the case where velocity information is available, the probability of an optical and $`\mathrm{H}\mathrm{I}`$ velocity coincidence within 2$`\sigma `$ of the $`\mathrm{H}\mathrm{I}`$ velocity dispersion of $``$13 $`\mathrm{km}\mathrm{s}^1`$ (Blitz et al. 1999) is about 0.06 within the velocity range in which HVCs are detected: about 800 $`\mathrm{km}\mathrm{s}^1`$. This probability must be increased somewhat because we consider only velocities outside the range of normal Galactic emission, but in the directions in which most dSphs are found, the Galactic emission is not wider than about 200 $`\mathrm{km}\mathrm{s}^1`$. The joint probability of velocity and spatial coincidence is thus $`10^3`$. Probabilities of chance coincidences can be estimated for individual galaxies and are done so where appropriate in the text. If emission was found to be associated with a galaxy, based on positional coincidence alone, or where possible, joint position and velocity coincidence, we averaged all of the profiles where emission was evident. One or more Gaussians were then fit to the averaged spectrum in order to determine a central velocity and velocity extent of the emission. The maps are shown in Figure 1. The values of the central velocity, FWHM velocity extent, and brightness temperature of the Gaussian fits to the position-averaged spectra are listed in Table 1. The derived values for M($`\mathrm{H}\mathrm{I}`$) and $`\mathrm{\Omega }`$, the $`\mathrm{H}\mathrm{I}`$ mass and the solid angle of the cloud, are not corrected for source convolution with the telescope beam. We then searched a 7$`\times `$7 area to see if the emission is localized around the target galaxy or whether it exists over a more extended area. In two cases, rather extended emission was found and the maps are shown in Figures 2 and 3. The individual maps are discussed in §3.1 below. Some of the apparent detections were quite weak and Jay Lockman kindly observed some of these with the NRAO 140 telescope<sup>1</sup><sup>1</sup>1The National Radio Astronomy Observatory is operated by Associated Universities, Inc., under cooperative agreement with the National Science Foundation. prior to its shutdown in late July 1999 to obtain confirmation. In most cases the observations were about 4 times longer than those of the LDS. The beam of the 140 telescope is about 20 at the frequency of the 21-cm line. Finally, we searched the catalogues of Hulsbosch & Wakker (1988) and Wakker & van Woerden (1991) to see whether any of the $`\mathrm{H}\mathrm{I}`$ clouds are catalogued HVCs. We found two such cases, the cloud associated with And V is HVC 368, and the cloud associated with Sculptor is HVC 561. LGS 3 had been previously detected at three positions and DDO 210 at one position on the 1$`\times `$1 sampling grid of Hulsbosch & Wakker (1988). The temperature-weighted mean Galactic latitude and longitude, as well as the $`\mathrm{H}\mathrm{I}`$ column density and mass for each cloud can be found in Table 1. The derived $`\mathrm{H}\mathrm{I}`$ masses assume that the emission is at the distance of the galaxy. The non-detections are listed in Table 2. The observed properties of the detected clouds are similar to one another and to the typical properties of HVCs, but are generally weaker than $`\mathrm{H}\mathrm{I}`$ detections of Local Group dIrr, Irr, or spiral galaxies. ## 3 Results We have made new $`\mathrm{H}\mathrm{I}`$ detections toward And III, And V, Leo I and Sextans. In addition, the $`\mathrm{H}\mathrm{I}`$ emission toward LGS 3, DDO 210, and Sculptor is found to be considerably more extended than previously thought. In all, of 21 confirmed Local Group dSph galaxies, $`\mathrm{H}\mathrm{I}`$ is detected in apparent association with 10 galaxies. Two galaxies are too far south to have been observed in the Leiden-Dwingeloo survey, and several probable non-detections are somewhat ambiguous. One galaxy previously reported as a detection, Antlia (Fouqué et al. 1990), is probably an instrumental artifact. There are 9 non-detections, not including the dE systems or the two unconfirmed candidate dSph galaxies of Karachentsev & Karachentseva (1999). Four galaxies are below the declination limit of the LDS, but two of these, Phoenix and Tucana, have been observed and detected in $`\mathrm{H}\mathrm{I}`$ by others (Young & Lo 1997; Oosterloo et al. 1996); they are therefore included in the list of detections. Thus almost half of the confirmed Local Group dSph galaxies have been detected in $`\mathrm{H}\mathrm{I}`$, though only five of those detected have measured optical velocities as of this writing. Attempts are actively underway by several groups to obtain more velocities. Two of the galaxies shown in Figure 1 exhibit emission close to, but not coincident with the target galaxy: Leo I and Sextans. This might explain why these two galaxies have not been previously detected in $`\mathrm{H}\mathrm{I}`$; and why the detection of $`\mathrm{H}\mathrm{I}`$ toward Tucana by Oosterloo, Da Costa, & Stavely-Smith (1996) ought to be reinterpreted. These authors found a cloud very close to, but not coincident with Tucana, in which the highest column density emission is offset by about 15′ from the nucleus. Oosterloo et al. (1996) felt that the position offset implied that the $`\mathrm{H}\mathrm{I}`$ cloud and the galaxy are unrelated. We show in the next section, however, that the relationship between the $`\mathrm{H}\mathrm{I}`$ cloud and Tucana is similar to other galaxies in our sample, particularly Leo I, in which the velocity of the $`\mathrm{H}\mathrm{I}`$ and the galaxy are both measured and found to be in close agreement. We therefore include Tucana in Table 1 as a detection. ### 3.1 Comments on Individual Detections #### 3.1.1 And III And III is a rather problematic $`\mathrm{H}\mathrm{I}`$ detection. It appears to be clearly detected in the LDS (see Figure 1), and a confirmation spectrum of the galaxy taken with the 140 telescope showed a much weaker component at low significance at the same velocity. The weakness of the 140 spectrum is rather surprising, however, given the smaller beam and the longer integration time. The detection, nevertheless, does appear to be real, though of low significance. A measurement of the optical velocity of the galaxy or a deep $`\mathrm{H}\mathrm{I}`$ integration over the LDS beam will confirm or refute this detection. The values given in Table 1 are from the deeper 140 observations. (NB: The referee, Mario Mateo, has pointed out that the optical velocity of this galaxy has recently been determined by Côté, Mateo, & Sargent (2000), and is within 1$`\sigma `$ of the combined $`\mathrm{H}\mathrm{I}`$ and optical velocity uncertainties). The $`\mathrm{H}\mathrm{I}`$ non-detection reported by Thuan & Martin (1979) is for a 4 beam at Arecibo centered on the galaxy. The $`\mathrm{H}\mathrm{I}`$ mass given in Table 1 is consistent with their upper limits, particularly if the $`\mathrm{H}\mathrm{I}`$ is not centered directly on the dSph, as is the case for several other galaxies. #### 3.1.2 And V This is an intriguing case not only because of the apparent association of the $`\mathrm{H}\mathrm{I}`$ with the galaxy, but also because the $`\mathrm{H}\mathrm{I}`$ is listed as HVC 368 in the compilation of Wakker & van Woerden (1991). Thus a measurement of the radial velocity of the galaxy that agrees with the $`\mathrm{H}\mathrm{I}`$ velocity would provide the second firm association of an HVC with a dSph (see the discussion of Sculptor below). As shown in Figure 2, however, HVC 368 lies quite close to HVC 287, also known as complex H. The latter has an angular extent of $``$205 deg<sup>2</sup>, and a velocity similar to HVC 368. Blitz et al. (1999) have argued that complex H lies beyond the $``$40 kpc radius of the $`\mathrm{H}\mathrm{I}`$ disk of the Galaxy, but probably not far beyond. The proximity of the two clouds and the similarity of their radial velocities suggests that the smaller cloud may be a fragment of the larger one, and may thus be an unrelated foreground object. In that case, the probability of a chance coincidence is much higher than the value of $`10^2`$ for the other dSph galaxies. We present additional arguments in §4 why this $`\mathrm{H}\mathrm{I}`$ cloud might be a chance superposition. A measurement of the optical velocity would clearly determine whether the $`\mathrm{H}\mathrm{I}`$ and the galaxy are associated. #### 3.1.3 DDO 210 The $`\mathrm{H}\mathrm{I}`$ in this galaxy was previously observed by Lo, Sargent, & Young (1993) at the VLA, and the mass we obtain for the $`\mathrm{H}\mathrm{I}`$ emission centered on the galaxy is in good agreement with theirs, correcting for the different distances assumed. Note, however the additional emission seen in Figure 1 that is outside the Lo et al. (1993) field of view. The relatively strong emission seen at $`l`$ = 35, $`b`$ = -32 is a narrow component at nearly the same velocity as the galaxy. The low level disconnected emission in Figure 1 may not be real, but the $`\mathrm{H}\mathrm{I}`$ cloud associated with DDO 210 nevertheless appears to be considerably larger than that observed by Lo et al. (1993). #### 3.1.4 Leo I There is a clear detection of an $`\mathrm{H}\mathrm{I}`$ cloud with a large velocity dispersion in the direction of this galaxy at a velocity close to, but slightly shifted from the velocity of the dSph. The uncertainty in the velocity of the line center is, however, relatively large and the width of the line comfortably encompasses the velocity of the galaxy. We show two maps in Figure 3. On the left is a map of the brightness temperature distribution over the full width of the $`\mathrm{H}\mathrm{I}`$ line. The map shows the extent of what appears to be highly fragmented $`\mathrm{H}\mathrm{I}`$ emission; a larger map (not shown) indicates that this emission is confined to the area shown in the panel. As a check, we produced a map over the same area within $`\pm `$10 $`\mathrm{km}\mathrm{s}^1`$ of the Leo I optical velocity shown in the right hand panel of Figure 3. The $`\mathrm{H}\mathrm{I}`$ in this velocity range is confined to an area quite close to the galaxy, but surprisingly, no $`\mathrm{H}\mathrm{I}`$ is seen directly toward the galaxy itself. The full velocity extent of the $`\mathrm{H}\mathrm{I}`$ is much more extended, and is seen over an area of about 10–20 deg<sup>2</sup>. Nevertheless, since the $`\mathrm{H}\mathrm{I}`$ in the velocity range of Leo I is so closely confined to the environs of the galaxy, it seems likely that all of the $`\mathrm{H}\mathrm{I}`$ is at the distance of the dSph. Higher resolution $`\mathrm{H}\mathrm{I}`$ mapping might make the association clearer. The $`\mathrm{H}\mathrm{I}`$ non-detection reported by Knapp, Kerr, & Bowers (1978) was made with a single pointing of the 300 ft (91 m) telescope at Green Bank and a 10 beam; it is consistent with the present detection. #### 3.1.5 LGS 3 The $`\mathrm{H}\mathrm{I}`$ cloud in the vicinity of LGS 3 was noted by Hulsbosch & Wakker (1988) and is quite close to the galaxy, but the velocity centroid of most of the $`\mathrm{H}\mathrm{I}`$ differs from that of the galaxy by about 50 $`\mathrm{km}\mathrm{s}^1`$. This galaxy was mapped with the VLA by Young & Lo (1997) who found an $`\mathrm{H}\mathrm{I}`$ cloud centered on the dSph within a few $`\mathrm{km}\mathrm{s}^1`$ of the velocity of LGS 3 and a total $`\mathrm{H}\mathrm{I}`$ mass of $`4\times 10^5`$ $`\mathrm{M}_{}`$. Smoothing of the LDS $`\mathrm{H}\mathrm{I}`$ profile at the position of the galaxy shows both the -285 $`\mathrm{km}\mathrm{s}^1`$ component and a component centered at about -340 $`\mathrm{km}\mathrm{s}^1`$. A hint of the more extreme velocity can be seen in the 140 spectrum shown by Young & Lo (1997). The probability of two HVCs seen along the same line of sight that are this compact is about 10<sup>-2</sup> (see §2). While the large velocity difference between the cloud shown in Figure 2 and the cloud mapped by Young & Lo (1997) is puzzling, the relation of the dSph to the cloud bears some resemblance to the other systems pictured in Figure 1. We discuss in §4 the possibility that the more negative velocity component may result from ram-pressure stripping. Although the emission detected by Young & Lo (1997) is not seen in either the map or the spectrum shown in Figure 1, both are consistent with their detection. The peak flux density of this galaxy observed at the 140 telescope is about 110 mJy, corresponding to an expected antenna temperature of 0.021 K when observed with the Dwingeloo telescope. This is the strength at which the feature is seen, within the noise, in the LDS at the position of the galaxy. The spectrum shown in Figure 1 is an average of 6 positions at which the galaxy is detected; the expected peak temperature of that feature is lowered by yet another factor of 2.5; a feature of that strength is too weak to be detected. Both the spectrum and the map of this source are consistent with the observations of Young & Lo (1997). #### 3.1.6 Pegasus Pegasus has been previously mapped by Lo, Sargent, & Young (1993). The mass obtained with the LDS data is in reasonable agreement with that of Lo et al. (1993) when account is taken of the different distances assumed and the possibility that some of the flux may be missed with the VLA observations. #### 3.1.7 Phoenix Although this galaxy is too far south for the LDS, it has previously been mapped by Carignan, Demers, & Côté (1991), Young & Lo (1997) and St-Germain et al. (1999). Because this galaxy lies close to the main emission from the Milky Way, and its optical velocity has not yet been measured, it is unclear which of two relatively compact velocity components, if either, is associated with the galaxy. The $`\mathrm{H}\mathrm{I}`$ properties listed in Table 1 are from St-Germain et al. (1999). They cite evidence suggesting that the lower mass (i.e. $`2\times 10^5`$ $`\mathrm{M}_{}`$) $`\mathrm{H}\mathrm{I}`$ component is associated with the galaxy, in part because they find the larger mass component implausibly large. We follow their suggestion and use the lower value in Figure 4 below. Without a reliable single dish map however, it is unclear whether the interferometer recovers all of the flux associated with the galaxy. Thus the mass quoted by St-Germain et al. (1999) should be considered a lower limit. #### 3.1.8 Sextans Although the emission from Sextans shown in Figure 1 is not centered on the galaxy, the velocity of the $`\mathrm{H}\mathrm{I}`$ emission is quite close to the optical velocity. The probability of a cloud outside the galaxy with a velocity within $`\pm `$15 $`\mathrm{km}\mathrm{s}^1`$ of the galaxy is about 0.05. The joint probability of a small cloud within 4 square degrees with a velocity outside of the normal Galactic emission in such close agreement with the optical velocity is about $`1.5\times 10^3`$. The cloud shown in Figure 1 is therefore probably associated with the galaxy; the edge of the $`\mathrm{H}\mathrm{I}`$ cloud is only about 1 kpc from the galaxy in projection. #### 3.1.9 Sculptor Sculptor is a particularly interesting case because of the two-lobed $`\mathrm{H}\mathrm{I}`$ structure found by Carignan et al. (1998) and because the $`\mathrm{H}\mathrm{I}`$ emission associated with the galaxy is catalogued as HVC 561 in the compilation of Wakker & van Woerden (1991). The map of the galaxy shown in Figure 1 is incomplete because the galaxy is beyond the southern declination limit of the LDS. Carignan et al. (1998) were at pains to point out that the $`\mathrm{H}\mathrm{I}`$ emission in their map might only be a small fraction of the total because of the spatial filtering of their interferometric observations, and because the $`\mathrm{H}\mathrm{I}`$ extends to the limit of the spatial scale to which their observations are sensitive. A more complete map was recently published by Carignan (1999) using the Parkes 43-m telescope showing that the $`\mathrm{H}\mathrm{I}`$ associated with Sculptor has an extent of about 2, larger than that shown in Figure 1, but the total mass is not given. The mass given in Table 1 is from the LDS, but may be low by a factor of about 2-3. #### 3.1.10 Tucana Tucana is below the southern declination limit for the LDS, but was observed at the ATNA by Oosterloo, Da Costa, & Staveley-Smith (1996), who found an $`\mathrm{H}\mathrm{I}`$ cloud close to, but not quite coincident with the galaxy. The centroid of the emission is located only 15 from the galaxy and some of the emission is as close as 1 from it. The joint probability of such close agreement in position and velocity between the observed $`\mathrm{H}\mathrm{I}`$ and the optical galaxy is about 10<sup>-4</sup>, suggesting that the $`\mathrm{H}\mathrm{I}`$ cloud is indeed bound to the galaxy. Furthermore the Oosterloo et al. (1996) observations were done with the ATNA, and the map extends over a large fraction of the beam. It is therefore likely that the extent of the $`\mathrm{H}\mathrm{I}`$ emission is larger than that shown in their map, similar to what is observed in Sculptor. The relationship of the $`\mathrm{H}\mathrm{I}`$ to the optical galaxy in this case is similar to galaxies such as Leo I, And V, Sculptor and possibly LGS 3. The low probability of a chance superposition, together with the similarity to other systems, suggests that the $`\mathrm{H}\mathrm{I}`$ is indeed related to Tucana. If the cloud is at the distance of the dSph, it has a mass of $`1.5\times 10^6`$ $`\mathrm{M}_{}`$. ### 3.2 The Non-Detections A number of the non-detections are somewhat ambiguous because the optical velocity of the galaxy is close to the Galactic $`\mathrm{H}\mathrm{I}`$ emission, or because a feature of marginal significance is present in the LDS. High significance non-detections were made toward And II, Leo II and And VII using the 140 telescope with an rms noise temperature of about 35 mK in 1 $`\mathrm{km}\mathrm{s}^1`$ channels. The non-detection toward Leo II is consistent with that of Young (1999). A reported detection of Antlia by Fouqué et al. (1990) could not be reproduced, and is probably part of the emission from NGC 3109 entering a sidelobe of the telescope. We were unable to see any evidence for emission in the LDS at the position observed by these authors. ## 4 Discussion The detection of $`\mathrm{H}\mathrm{I}`$ associated with many dSph galaxies suggests that rather than being gas-free systems, perhaps half of them are in fact gas-rich galaxies, with $`\mathrm{H}\mathrm{I}`$ that is a substantial fraction of the luminous mass, M<sub>L</sub>, and in some cases exceeding it. If we take M<sub>L</sub>/L<sub>V</sub> = 1-2 in solar units, then more than half of the galaxies in Table 1, about 25% of all of the dSphs, have $`\mathrm{H}\mathrm{I}`$ masses in excess of 50% of the luminous mass of the galaxy. Most previous observations have concentrated on searching for $`\mathrm{H}\mathrm{I}`$ only at the central position of a galaxy largely because the beam of the $`\mathrm{H}\mathrm{I}`$ observations was as large or larger than the galaxy itself. The observations presented here show that sometimes the $`\mathrm{H}\mathrm{I}`$ avoids the nucleus of the galaxy and is seen over a large area around it (e.g. Leo I). With the exception of And V, there is little doubt that the $`\mathrm{H}\mathrm{I}`$ clouds shown in the maps in Figure 1 are associated with the galaxies, even though optical radial velocities are not available for most dSphs. As discussed in §2, the probability of a chance superposition with an $`\mathrm{H}\mathrm{I}`$ cloud is quite small, typically about 0.01, and if a velocity is available, the probability is $`<10^3`$. If all of the galaxies shown in Table 1 have ionized hydrogen masses equal to that of their atomic hydrogen, and if the $`\mathrm{H}\mathrm{II}`$ were smoothly distributed out to the same radius as the $`\mathrm{H}\mathrm{I}`$, the typical emission measure except for Phoenix and Tucana would be 10<sup>-3</sup> – 10<sup>-4</sup> cm<sup>-6</sup> pc, more than an order of magnitude below what is currently detectable. Thus the dSphs might harbor substantial quantities of ionized gas in addition to what is shown in Figure 1. We might reasonably ask why some of the dSph galaxies have large, massive $`\mathrm{H}\mathrm{I}`$ envelopes (such as Leo I), why some have relatively wimpy $`\mathrm{H}\mathrm{I}`$ envelopes (such as Sextans and Sculptor), and why some seem to be devoid of $`\mathrm{H}\mathrm{I}`$ entirely (such as Ursa Minor and Draco). To approach this question, we have plotted the the $`\mathrm{H}\mathrm{I}`$ mass of each of the dSph and dIrr galaxies in the Local Group as a function of the distance from the nearest giant spiral, either M31 or the Milky Way (MW) in Figure 4a. The data for the dIrr galaxies are taken from Mateo (1998). Figure 4 shows a rather remarkable effect. With the exception of And V, no galaxy within 250 kpc of either M31 or the MW has an $`\mathrm{H}\mathrm{I}`$ mass in excess of 10<sup>6</sup> $`\mathrm{M}_{}`$; beyond 250 kpc almost all of the dwarf galaxies have substantial $`\mathrm{H}\mathrm{I}`$ envelopes in excess of 10<sup>6</sup> $`\mathrm{M}_{}`$, regardless of whether the galaxies are dSph or dIrr. Inside the 250 kpc cutoff most of the upper limits and the three detections are in fact below 10<sup>5</sup> $`\mathrm{M}_{}`$. The one exception is And V which is the galaxy most likely to be a chance superposition (see §3.1.2). Antlia and Phoenix are the exceptions beyond the 250 kpc cutoff. Antlia may be anomalous because of its close proximity to NGC 3109. The $`\mathrm{H}\mathrm{I}`$ mass of Phoenix may be underestimated since its mass is determined from an interferometric measurement confused by bright foreground emission at the same velocity. It is also possible that another more massive $`\mathrm{H}\mathrm{I}`$ component is actually associated with the galaxy (§4.1.7). To check whether the effect seen in Figure 4a might be a distance effect related to the Malmquist bias, we replotted the figure by normalizing to the visual luminosity of each galaxy, L<sub>V</sub>, taken from Mateo (1998); the results are shown in Figure 4b. Evidence for the 250 kpc $`\mathrm{H}\mathrm{I}`$ cutoff remains quite strong in this plot. Van den Bergh (1999a) noted that the dSph galaxies tend to be closer to M31 and the Milky Way than the dIrr galaxies, but because some of the galaxies beyond the 250 kpc $`\mathrm{H}\mathrm{I}`$ cutoff are dSphs, the segregation in $`\mathrm{H}\mathrm{I}`$ properties is not simply a function of morphological type. The sharp cutoff implies that some process strips the galaxies with perigalacticons $`<`$ 250 kpc of their $`\mathrm{H}\mathrm{I}`$, a suggestion first made by Einasto et al. (1974) and later by Lin & Faber (1983). We investigate the relative importance of ram-pressure stripping and tidal stripping below. ### 4.1 Ram-Pressure Stripping If the sharp boundary at 250 kpc is due to ram-pressure stripping by hot halo gas (Gunn & Gott 1972), one can estimate the density of the ambient gas responsible for the stripping, $`\rho _{}`$, from $$\rho _{}>\alpha G\mathrm{\Sigma }_{}\mathrm{\Sigma }_g/V^2,$$ (1) where $`\mathrm{\Sigma }_{}`$ and $`\mathrm{\Sigma }_g`$ are the the stellar and gas surface densities respectively in the dwarf galaxy, $`V`$ is the velocity of the galaxy through the hot halo gas, and $`\alpha `$ is a constant near unity that depends on whether the galaxy is a flattened or a spheroidal system, and on the functional form of the gravitational potential. For a uniform spherical stellar system with an extended uniform density halo, $`\alpha =\pi /6`$. $`\mathrm{\Sigma }_{}`$ and $`\mathrm{\Sigma }_g`$ in this case are the peak values measured at the center of the galaxy. The inequality exists because the gas may be clumped and the surface filling fraction of the $`\mathrm{H}\mathrm{I}`$ may be less than unity, raising the local value of $`\mathrm{\Sigma }_g`$ above its beam averaged value. If the gravitational mass of the galaxy, M, is given by $`Rv_{3\mathrm{D}}^2/G`$, then we may rewrite Equation 1 as: $$\rho _{}>\frac{\mathrm{\Sigma }_gv_{3\mathrm{D}}^2}{4RV^2}.$$ (2) In terms of observables, Equation 2 may be rewritten as $$n_{}>1.12\times 10^4\frac{T_B(\mathrm{\Delta }V)^3}{RV^2}\mathrm{cm}^3,$$ (3) where $`n_{}`$ is the number density of the hot halo gas, $`T_B`$ is the peak brightness temperature of the $`\mathrm{H}\mathrm{I}`$, and $`\mathrm{\Delta }V`$ is the measured full width at half maximum of the $`\mathrm{H}\mathrm{I}`$ line. We use this equation to estimate $`n_{}`$. For $`V`$ we take the one-dimensional velocity dispersion of 60 $`\mathrm{km}\mathrm{s}^1`$ for the Local Group dwarf galaxies (van den Bergh 1999a). For $`\mathrm{\Delta }V`$, we take the mean value for the galaxies from Table 1 of 25 $`\mathrm{km}\mathrm{s}^1`$. For $`R`$, we take it equal to $`R_g`$, the radius to which gas is observed in a dwarf galaxy; it ranges from about 1 - 10 kpc for all of the galaxies in the sample with the larger values all outside the 250 kpc cutoff radius. Inside the cutoff radius, three galaxies, And III, Sculptor and Sextans, still have associated $`\mathrm{H}\mathrm{I}`$, and we take $`R=2`$ kpc, the largest of the radii to which gas is observed. Thus $`n_{}2.4\times 10^5`$ $`\mathrm{cm}^3`$. Interestingly, this value is close to the value of $`5\times 10^5`$ $`\mathrm{cm}^3`$ derived by Moore & Davis (1994) for ram-pressure stripping of the Magellanic Clouds. Their value applies to the density of a hot halo at a distance of 65 kpc from the Galactic Center, whereas ours is an average over a volume of 250 kpc radius. If the hot gas is in virial equilibrium and the mass of the MW is 1 – $`1.5\times 10^{12}`$ $`\mathrm{M}_{}`$, the hot gas temperature is $``$ 1 – $`1.4\times 10^6`$ K, and the thermal pressure $`P/k`$ = 23 - 34 K $`\mathrm{cm}^3`$. The total mass of the hot halo at the derived $`n_{}`$ is about $`1\times 10^{10}`$ $`\mathrm{M}_{}`$ to 250 kpc. Gas will be stripped from the galaxy if the column density of hot gas is equal to the column density in the galaxy. Thus for galaxies (or HVCs) with N($`\mathrm{H}\mathrm{I}`$) $`>n_{}\times `$ 250 kpc = $`1.8\times 10^{19}`$ $`\mathrm{cm}^2`$, the gas can remain in the galaxy until it collides with the MW. The most recent map of Sculptor by Carignan (1999) shows $`\mathrm{H}\mathrm{I}`$ contours ranging from 0.15 - $`2.4\times 10^{19}`$ $`\mathrm{cm}^2`$, but with a broad emission plateau of about $`0.6\times 10^{19}`$ $`\mathrm{cm}^2`$, somewhat below the ram-pressure limit. The discrepancy, while not large, may be due to clumping (implying higher mean column densities within the beam), a somewhat high estimate for $`n_{}`$, Sextans not having traversed a full 250 kpc in its orbit through the hot halo, or some combination of the three. The radial velocity of Sculptor relative to the Galactic Standard of Rest (GSR) is 75 $`\mathrm{km}\mathrm{s}^1`$, close to the value of 60 $`\mathrm{km}\mathrm{s}^1`$ assumed. ### 4.2 Tidal Stripping Gas will be tidally stripped from a galaxy if $$\frac{GM}{R^2}>\frac{\mathrm{d}}{\mathrm{d}r}\left(\frac{\mathrm{\Theta }^2}{r}\right)R,$$ (4) where $`r`$ is the distance from the MW or M31 to the galaxy, and $`\mathrm{\Theta }`$ is the circular speed of the MW or M31 at the distance of the dwarf galaxy. At a distance of 250 kpc, $`\mathrm{\Theta }`$ = 130 $`\mathrm{km}\mathrm{s}^1`$ (appropriate for an enclosed MW mass of $`1\times 10^{12}`$ $`\mathrm{M}_{}`$); the tidal radius of a dwarf galaxy with a total mass of $`1\times 10^8`$ $`\mathrm{M}_{}`$ is about 11.5 kpc, greater than the largest radius to which $`\mathrm{H}\mathrm{I}`$ is detected in any Local Group dwarf. Thus for these parameters, the gas is tidally stable. For dwarf galaxies at 80 kpc from the MW, the distance of Sculptor and Sextans, the tidal radius is about 4 kpc, comfortably larger than the 1.5 – 2 kpc to which $`\mathrm{H}\mathrm{I}`$ is detected in either Sculptor or Sextans. Thus, the gas in these galaxies is tidally stable at their current distance, but if their perigalacticon is significantly smaller, then even the present day observed gas could be tidally stripped from these galaxies. Thus tidal stripping can be important in removing the outer $`\mathrm{H}\mathrm{I}`$ envelopes of the dSph galaxies within 250 kpc of the MW and M31. Nevertheless, most of the dSphs within 250 kpc are devoid of $`\mathrm{H}\mathrm{I}`$ to the current limits of detectability. While Galactic tides can be important in stripping the outermost $`\mathrm{H}\mathrm{I}`$ layers of Local Group dSphs, it is difficult to understand how tidal stripping can completely rid a dSph of its atomic gas. It is instructive to compare the effect of tidal vs. ram-pressure stripping in a dwarf galaxy at a radius of 10 kpc and located at a distance of 250 kpc from the MW. The tidal acceleration from Equation 4 above is $`R\mathrm{\Theta }^2/r^2`$. The acceleration of a parcel of gas due to ram-pressure is $$\frac{\rho _{}V^2}{\mathrm{\Sigma }_g}=\frac{\rho _{}V^2}{\mathrm{N}(\mathrm{H}\mathrm{I})\mu m_\mathrm{H}},$$ (5) where $`\mu `$ is the mean mass per nucleon of the atomic gas. For $`\mathrm{\Theta }=130`$ $`\mathrm{km}\mathrm{s}^1`$, $`n_{}`$ = $`2.5\times 10^5`$ $`\mathrm{cm}^3`$, and N($`\mathrm{H}\mathrm{I}`$) = $`5\times 10^{18}`$ $`\mathrm{cm}^2`$, a typical value in the outskirts of Leo I and LGS 3, the ratio of the ram-pressure acceleration to the tidal acceleration is about 18. Tidal stripping becomes relatively more important, however, as one gets closer to the center of the Local Group giant spirals, but if there is even one tenth the density of hot gas around the MW and M31, as we derive above, ram-pressure stripping will be relatively more important than tidal stripping unless perigalacticon is very small. We therefore conclude that although tidal stripping may play a role in the gas depletion of the dwarf galaxies in the Local Group, tidal stripping alone cannot explain the near total absence of atomic gas in most of the dSphs within 250 kpc of either the MW or M31. Ram-pressure stripping is much more efficient over most of the parameter space permitted by the $`\mathrm{H}\mathrm{I}`$ observations, and implies that the MW and M31 have hot halos with radii of $``$250 kpc, beyond which stripping is ineffective. It is quite reasonable that the MW and M31 have hot halos in the context of the Blitz et al. (1999) picture of the formation of the Milky Way and M31 from the accretion of HVCs. The dynamical model presented by these authors is quite simple and does not include collisions between HVCs which are expected close to either galaxy or in the region between the galaxies. Such collisions are, however, expected; the typical center-of-mass collision velocity is about 100 $`\mathrm{km}\mathrm{s}^1`$ at the present epoch, high enough to completely ionize the colliding clouds and to raise their temperatures to $`10^5`$ K (McKee & Hollenbach 1980). Collisions between clouds at the present epoch are, however, expected to be rare, but should have been more frequent in the early universe (see Fig. 16 in Blitz et al. 1999). The typical cloud collision velocity at those times would have been higher, leading to hot gas temperatures which can reach the virial values. The cooling time of the hot halo gas at the present epoch is $`3\times 10^{10}`$ yr at an inferred mean density of $`2.5\times 10^5`$ $`\mathrm{cm}^3`$. Thus the hot halo gas is stable for more than a Hubble time, but the inner parts can cool and condense in a shorter time if the density distribution is isothermal. Whether HVC collisions are frequent enough to produce a hot halo can be tested by direct numerical simulations. The halos around each galaxy need not be separate entities and may be connected along the line between the two galaxies. The existence of a hot halo is consistent with the destruction of $``$1000 objects with $`\mathrm{H}\mathrm{I}`$ content of $`10^7`$ $`\mathrm{M}_{}`$, close to the typical mass derived by Blitz et al. (1999) for the HVCs. It is worth noting that Klypin et al. (1999) have shown that simulations of hierarchical structure formation seem to require about 1000 dwarf galaxies to have formed in the Local Group, one and a half orders of magnitude more than have currently been identified. Klypin et al. (1999) suggested that the required number of dwarfs might be consistent with the inferred population of HVCs, but it is also possible that the gaseous halo is from a population of true dwarf galaxies with a typical $`\mathrm{H}\mathrm{I}`$ content similar to that found in the present-day dwarfs that have been destroyed by collisions. A mass of $`10^{10}`$ $`\mathrm{M}_{}`$ of hot gas would be provided by about $`10^210^4`$ dwarf galaxies, with typical $`\mathrm{H}\mathrm{I}`$ masses between $`10^610^8`$, the range of $`\mathrm{H}\mathrm{I}`$ masses in present-day dwarfs not yet stripped of $`\mathrm{H}\mathrm{I}`$. The mean free path for collisions between HVCs ought to be smaller than that for dwarfs because of their larger diameters, and the difference between the two cases might be testable by simulations. ### 4.3 The Velocity Anomalies of LGS 3 and Leo I One of the more difficult of the $`\mathrm{H}\mathrm{I}`$ observations to understand is that of LGS 3, and to a lesser extent, Leo I. LGS 3 has an $`\mathrm{H}\mathrm{I}`$ cloud clearly associated with it, but the $`\mathrm{H}\mathrm{I}`$ cloud differs in velocity from the galaxy by 50 $`\mathrm{km}\mathrm{s}^1`$. The observations of this galaxy by Young & Lo (1997) show a well-defined $`\mathrm{H}\mathrm{I}`$ component centered on the galaxy with a velocity essentially identical to the systemic velocity of the stars. The extent of the gas centered on the galaxy is only about 5′ compared to the extent of about 1$`\text{.}^{}`$5 in Figure 1. With a velocity difference of 50 $`\mathrm{km}\mathrm{s}^1`$, the gas shown in Figure 1 cannot be gravitationally bound to LGS 3, which suggests a chance superposition. On the other hand, the probability of a chance spatial superposition is about 1%, and the probability that the velocity would be so close to the systemic velocity is about 0.2, for a joint probability of about $`2\times 10^3`$. The situation with Leo I is similar but less extreme. The right hand panel of Figure 3 shows that $`\mathrm{H}\mathrm{I}`$ at the velocity of the galaxy is closely associated with it, even though no $`\mathrm{H}\mathrm{I}`$ is detected toward the stars themselves. Nevertheless, the velocity centroid of the emission shown in the right hand panel of Figure 3 differs from the systemic velocity of the galaxy by about 30 $`\mathrm{km}\mathrm{s}^1`$. In this case, however, the gas shown in both panels is part of the same velocity component. While the extended emission shown in Figure 3 is unlikely to be gravitationally bound to Leo I, it is more difficult to argue that this gas is also a chance superposition, not only because the lower velocity gas is part of the same velocity component, but also because the joint probability of having two galaxies in this small sample with two $`\mathrm{H}\mathrm{I}`$ clouds in the same line of sight at velocities close to the systemic velocity is $`<`$ 10<sup>-5</sup>. We note, however, that both galaxies are close to the 250 kpc boundary shown in Figure 4. Leo I is 250 kpc from the Milky Way (Mateo 1998) and LGS 3 is 270 kpc from the center of M31. Could it be that the $`\mathrm{H}\mathrm{I}`$ seen in Figure 1 for each galaxy is beginning to be stripped by the ram-pressure of the hot halo gas? Certainly for LGS 3, the gas in Figure 1 looks as if it has been swept away from the galaxy itself. Tidal stripping is not an option for either galaxy because there is a systematic offset in velocity in only one sense; tidal stripping would produce plumes with velocities both larger and smaller than the systemic velocity of the dSph. From Equation 5, we find that the mean acceleration of the gas due to tidal stripping is $`1.3\times 10^{10}`$ cm s<sup>-2</sup>. Gas is detected as much as 1$`\text{.}^{}`$5 from LGS 3, or about 20 kpc in projection from the galaxy. With a velocity difference of 50 $`\mathrm{km}\mathrm{s}^1`$, gas farthest from the galaxy will have taken $`4\times 10^8`$ yr to reach that distance. Assuming that the acceleration is constant, ram-pressure stripped gas would attain a velocity of $``$15 $`\mathrm{km}\mathrm{s}^1`$ in that time. This is a bit on the low side, but the agreement is not unreasonable, given the uncertainty with which the critical parameters are known. The density of the hot gas might be somewhat higher, as might the velocity of the galaxy with respect to the hot halo gas, and projection effects can be important at the 50% level. If the position and velocity offsets of the $`\mathrm{H}\mathrm{I}`$ from LGS 3 are due to ram-pressure stripping, one would expect there to be a velocity gradient in the sense that the most extreme differences in velocity are seen farthest from the galaxy. The individual $`\mathrm{H}\mathrm{I}`$ spectra that compose Figure 1 do not, however, show any detectable gradient in the extended cloud. Ram-pressure stripping should produce velocity differences from the systemic velocity with well-determined signs: the velocity of the stripped gas should always be closer to the systemic velocity of either the MW or M31 than that of the dwarf being stripped. In the case of Leo I, the galaxy has a GSR velocity of +178 $`\mathrm{km}\mathrm{s}^1`$; the anomalous velocity gas has a GSR velocity of +158 $`\mathrm{km}\mathrm{s}^1`$, in accord with expectations. For LGS 3, the velocity relative to the GSR is about 25 $`\mathrm{km}\mathrm{s}^1`$ more negative than that of M31, suggesting that the stripped gas should have a more positive velocity. But this is not the case; the stripped gas is 50 $`\mathrm{km}\mathrm{s}^1`$ more negative. This discrepancy is difficult to reconcile with ram-pressure stripping. We conclude that ram-pressure stripping can plausibly produce the magnitude of the velocity and positional offsets for LGS 3 and Leo I if the velocity of the galaxy relative to the hot gas and the density of the hot gas are within factors of two and three respectively of the values assumed for them. The sign of the velocity offset for LGS 3, however, seems inconsistent with ram-pressure stripping. ### 4.4 Comparison of Dwarf Galaxy Envelopes with HVC Properties In their paper on HVCs, Blitz et al. (1999) derived properties of HVCs under the assumption that they are typically at a distance of 1 Mpc; however they did not correct for beam smearing, which would lower both the diameters and the derived masses. Furthermore, if the typical HVC has a distance more like that of M31 than the 1 Mpc assumed, their derived masses may be too high by as much as a factor of 4, and their diameters too high by a factor of 2. A complete northern hemisphere HVC catalogue by Robishaw & Blitz (in preparation) will correct for beam smearing. Given this range of uncertainty, typical HVC $`\mathrm{H}\mathrm{I}`$ masses are about 5 - $`20\times 10^6`$ $`\mathrm{M}_{}`$ if the HVCs are extragalactic and typical diameters are about 15 - 28 kpc. Some of the dwarf galaxy $`\mathrm{H}\mathrm{I}`$ envelopes are just in this range, notably DDO 210, Leo I, LGS 3, IC 10, Leo A, Sextans A, and possibly And V. Several others are confused with the velocities from Galactic foreground emission, and although their masses are in the correct range, the full extent of their $`\mathrm{H}\mathrm{I}`$ emission is poorly determined. Thus, numerous Local Group dwarfs have $`\mathrm{H}\mathrm{I}`$ properties virtually indistinguishable from extragalactic HVCs, and as pointed out in §4, two previously catalogued HVCs are apparently associated with galaxies. This suggests that some of the HVCs might harbor undetected low surface brightness (LSB) galaxies, and searches are currently underway to detect galaxies toward the HVCs. If successful, this would bridge the gap between the non-detections of HVC analogues without stars in deep extragalactic $`\mathrm{H}\mathrm{I}`$ searches (Zwaan et al. 1997; Spitzak & Schneider 1998). On the other hand, the extragalactic $`\mathrm{H}\mathrm{I}`$ searches have sensitivities that trail off just at or above the derived typical HVC mass, especially if the masses derived by Robishaw & Blitz (in preparation) turn out to be lower, as expected. So there may nevertheless be a substantial population of low-mass intergalactic $`\mathrm{H}\mathrm{I}`$ clouds without associated stars. In either case, the HVCs might then turn out to be the missing dwarf galaxies in the simulations of Klypin et al. (1999). ### 4.5 Implications for Galaxy Formation and Evolution Mateo (1998), Grebel (1999) and van den Bergh (1999b) discuss several problems associated with the apparent lack of interstellar gas in the dSph galaxies, most notably, their complex star formation histories. Some galaxies seem to have had several episodes of star formation, including some as recently as 1 – 2 Gyr ago (Grebel 1999), but this hardly seems possible without at least some traces of gas that could have fueled this activity. The $`\mathrm{H}\mathrm{I}`$ observations presented in this paper suggest that one way around this problem is that all of the Local Group dwarf galaxies have had loosely bound $`\mathrm{H}\mathrm{I}`$ envelopes such as those seen for the galaxies at distances beyond the 250 kpc cutoff radius when the last episode of star formation took place. This loosely bound gas would be subject to small relatively localized perturbations that could lead either to star formation or gas disruption and may be why the star formation histories of the Local Group dwarfs are so heterogeneous (Mateo 1998; Grebel 1999). Dissipation in the gas might have generated both low levels of ongoing star formation as well as occasional large bursts until the orbits of the galaxies brought them within the hot halos around the MW and M31. Star formation would then have ended for the galaxies with $`\mathrm{H}\mathrm{I}`$ column densities insufficient to withstand the ram-pressure stripping. Clearly, the extended $`\mathrm{H}\mathrm{I}`$ envelope around Leo I is plausibly the source of the relatively recent star formation activity in that galaxy (e.g. Grebel 1998). It is perhaps only when the galaxies venture within a radius of 250 kpc that the dwarfs become stripped of their $`\mathrm{H}\mathrm{I}`$, eventually losing their ability to form stars. The dynamical simulations of HVCs in the context of the evolution of the Local Group (Blitz et al. 1999) suggest that at least some of the dwarf galaxies inside the 250 kpc cutoff may be approaching the MW and M31 for the first time. To traverse the entire length of the hot halo for galaxies at a velocity of 60 $`\mathrm{km}\mathrm{s}^1`$ appropriate for the Local Group dwarfs (van den Bergh 1999a) takes $`8\times 10^9`$ yr, a substantial fraction of a Hubble time. Thus some of the dwarfs such as Carina may only recently have lost their gas, while others such as Ursa Minor, that show no evidence of recent star formation activity, may have orbits that kept them within the cutoff distance for most of a Hubble time. If the connection between LSB dSph galaxies and the HVCs can be confirmed, it might also solve the metallicity problem for HVCs. That is, if the HVCs are extragalactic, they must have existed for a Hubble time, but the abundances and metallicities, while low, are not primordial (Wakker & van Woerden 1997; Wakker et al. 1999). How then did these HVCs get their metals? If the HVCs are associated with low levels of star formation as in the dSph galaxies, then the stars themselves could have contaminated the gas. With a range of \[Fe/H\] $``$ -1 to -2 dex for the stars in dwarf galaxies, one expects the gas to reflect these values. The best determined metallicity toward an HVC (complex C) is the measurement of S/H = 0.09 times the solar value (Wakker et al. 1999), as expected, but more measurements are needed. ## 5 Summary We have shown that there is good evidence that nearly half of the dwarf spheroidal galaxies in the Local Group contain large quantities of gas in an extended distribution around each galaxy. Even without measured velocities for most of the galaxies, the associations of the gas with the galaxies must be real because the number of positional coincidences is more than two orders of magnitude greater than would be expected from random placements of both the galaxies and the HVCs. The properties of the $`\mathrm{H}\mathrm{I}`$ associated with many of the dwarfs are similar to those expected for extragalactic HVCs, suggesting that the HVCs may harbor LSB galaxies similar to the dwarf galaxies in the Local Group. We have investigated the reason for the great diversity in the neutral gas content of Local Group dwarf galaxies, and have shown that within 250 kpc of the center of both the MW and M31, the gas content drops precipitously. Both tidal and ram-pressure stripping can play a role in removing the gas, but ram-pressure stripping is more effective and can strip a galaxy completely. The inferred mean density of hot gas is $`2.5\times 10^5`$ $`\mathrm{cm}^3`$. Two of the Local Group dwarfs, LGS 3 and Leo I, have $`\mathrm{H}\mathrm{I}`$ envelopes that differ from the systemic velocities by 50 $`\mathrm{km}\mathrm{s}^1`$ and 30 $`\mathrm{km}\mathrm{s}^1`$ respectively. The joint probability that both galaxies have unrelated clouds along the line of sight is $``$ 10<sup>-5</sup>. Ram-pressure stripping could cause large velocity offsets, and in the case of Leo I, the inferred hot halo properties are in reasonable quantitative agreement with what is observed. For LGS 3, the agreement is less good, but may still be within the range of acceptability. We find that the $`\mathrm{H}\mathrm{I}`$ observations can explain qualitatively the diversity in the star formation histories of the Local Group dwarfs, both the presence and absence of recent star formation in individual dSphs. We are grateful to Jay Lockman for obtaining $`\mathrm{H}\mathrm{I}`$ spectra at the 140′ telescope to help confirm some of the Leiden-Dwingeloo results and Tom Oosterloo for providing data on the Tucana dwarf. We would like to thank Taft Armandroff, Julianne Dalcanton, Eva Grebel, Raja Guhatakurtha, Dave Hollenbach, Chris McKee, David Spergel, Hy Spinrad, Amiel Sternberg, and Sidney van den Bergh for useful and lively discussions, some of which have been conducted electronically. The referee, Mario Mateo, made numerous useful suggestions, and provided the velocity of And III prior to publication.
warning/0001/cond-mat0001172.html
ar5iv
text
# Fluctuations and correlations in population models with age structure \[ ## Abstract We study the population profile in a simple discrete time model of population dynamics. Our model, which is closely related to certain “bit–string” models of evolution, incorporates competition for resources via a population dependent death probability, as well as a variable reproduction probability for each individual as a function of age. We first solve for the steady–state of the model in mean field theory, before developing analytic techniques to compute Gaussian fluctuation corrections around the mean field fixed point. Our computations are found to be in good agreement with Monte–Carlo simulations. Finally we discuss how similar methods may be applied to fluctuations in continuous time population models. \] The problem of population dynamics has attracted enormous interest over many years (for some introductions and recent applications see Refs. ). Beginning with simple logistic growth models , a tremendous variety of systems have been studied displaying a diverse range of behavior, varying from stable fixed points to strange attractors. Broadly speaking these models split naturally into two categories: those using continuous and those using discrete time. The simplest discrete time models describe species where there is no overlap between successive generations, leading to difference equations of the form $`N(t+1)=g[N(t)]`$, where $`N(t)`$ is the total population at time $`t`$. However these models may easily be generalized to species with multiple discrete age generations (for example: eggs, larvae, adults), where one or more generations may be present simultaneously. Instead of a single variable $`N`$, information about the age distribution is now carried in a “vector” $`𝐧(t)\{n_0(t),n_1(t),\mathrm{},n_D(t)\}`$, where $`n_a(t)`$ is the number of individuals of age $`a`$ at time $`t`$. Note that $`D`$ is the maximum age, while $`n_0`$ stands for the number of “newborns”. Beginning with the pioneering work of Leslie models of this type have been extensively analyzed. The simplest Leslie model is linear in $`𝐧`$, so that the evolution equation is just $`𝐧(t+1)=𝔸𝐧(t)`$. Here, $`𝔸`$ is the Leslie matrix $$𝔸=\left(\begin{array}{ccccc}f_0& f_1& \mathrm{}& & f_D\\ v_0& 0& \mathrm{}& & 0\\ 0& v_1& & & 0\\ \mathrm{}& & \mathrm{}& & \mathrm{}\\ 0& \mathrm{}& 0& v_{D1}& 0\end{array}\right),$$ (1) where the elements $`f_a`$ are the fecundities (number of offspring produced) of individuals of age $`a`$, and the $`v_a`$ are Verhulst factors (the fraction of individuals of age $`a`$ who survive to become age $`a+1`$). Though the original Leslie model had all the $`\{f_a\}`$ and $`\{v_a\}`$ as constants, generalizations have since been made to $`𝐧`$–dependent factors , so that the evolution dynamics becomes inherently non–linear. For example, $`𝐧`$–dependent Verhulst factors are often used to mimic competition for finite resources. However one deficiency of the models discussed hitherto is that they are deterministic. Real population systems are of course affected by random fluctuations, coming from the environment and/or from the intrinsic dynamics of the birth/death processes. Such stochastic Leslie models have also been investigated , however only for cases where the birth/death probabilities were independent of the population vector $`𝐧`$. To date no information has been available regarding the more realistic case of fluctuations in stochastic age structured models with population dependent birth/death probabilities . This is the situation we will study in this letter. Population models have also been intensively studied by physicists in recent years in the context of so–called “bit–string” models of evolution . These models are based on the mutation accumulation hypothesis , which assumes that during the aging process each individual accumulates exclusively late–acting deleterious genetic mutations. In “bit–string” models the genome of a particular species is encoded as a series of ‘$`0`$’s and ‘$`1`$’s (deleterious mutations), and as an individual ages the bits (genes) are activated one by one. When the accumulated sum of bad genes reaches a certain number the individual dies (although death may also occur at younger ages due to Verhulst competition). Note that the “bit–strings” of the offspring may differ from those of the parent due to additional beneficial (‘$`1`$$``$$`0`$’) or deleterious (‘$`0`$$``$$`1`$’) mutations. This type of model is clearly well suited to efficient computer simulation . In its simplest case, where individuals die after the first deleterious mutation, “bit–string” models simply correspond to multiple genome population models with age structure, where the different genomes can be distinguished by different maximum ages. Deterministic versions of some of these models have already been treated analytically . However, an analysis of the important role played by fluctuations has so far been lacking. Our calculations form the first step towards filling this gap. We begin our analysis by defining our discrete time population model. For simplicity, we consider only a single species reproducing asexually, without mutations. Thus, our system can be described by a single vector $`𝐧`$. At each time step we compute the Verhulst factor $`V(𝐧)`$ and let each individual survive with probability $`V`$. After this “pruning”, each of the remaining individuals of age $`a`$ may give birth to $`F_a`$ offspring with probability $`r_a`$. At this point the remaining population is aged by one time step, with the exception of the new offspring who make up $`n_0`$. Individuals who exceed the maximum age $`D`$ then die immediately and are removed from the system. Since our model allows for a variable reproduction probability $`r_a`$ as a function of age, features such as puberty and menopause can be naturally incorporated. However, we do assume that reproductive individuals of the same age produce an identical number ($`F_a`$) of offsprings. Note that we are not restricting ourselves to specific forms for $`V`$, $`r_a`$, or $`F_a`$ beyond some general features, so that this analysis may easily be applied to the various “bit–string” models . These general features include $`V,r_a[0,1]`$, since they represent probabilities. Furthermore, we assume that $`V`$ depends only on the total population $`N_an_a`$, via the ratio $`N/N_0`$, where $`N_0`$ represents a characteristic population size that the resources can support. Also, to be reasonable, $`V`$ is assumed to be monotonically decreasing with $`N`$. For comparisons with simulations, we use $`V=s_0\mathrm{exp}(N/N_0)`$, where $`s_0`$ is another constant, a form frequently chosen in the biology literature . In contrast, the algebraic form $`V=1(N/N_0)`$ is the favorite in the recent physics literature. We prefer the exponential form, since absolute cut-offs seem unrealistic in a real population system. As mentioned above, deterministic versions of this model have been studied ; in particular the population dynamics of a semivoltine species studied in Ref. is quite similar to a $`D=1`$ version of our model. Our goal is to go beyond these deterministic treatments and analyze the fluctuations and correlations in this system. Therefore, we need to consider $`P(𝐧,t)`$, the probability of finding the population with a particular distribution $`𝐧`$ at time $`t`$. Its evolution obeys the master equation $`P(𝐧,t+1)={\displaystyle \underset{m_0,\mathrm{},m_D,n_{D+1}}{}}P(𝐦,t)\times `$ (2) $`\left[{\displaystyle \underset{a=1}{\overset{D+1}{}}}\left({\displaystyle \genfrac{}{}{0pt}{}{m_{a1}}{n_a}}\right)V^{n_a}\left[1V\right]^{m_{a1}n_a}\right]{\displaystyle \underset{b_0,\mathrm{},b_D}{}}\delta \left[n_0{\displaystyle \underset{c=0}{\overset{D}{}}}F_cb_c\right]`$ (3) $`\times \left[{\displaystyle \underset{a=1}{\overset{D+1}{}}}\left({\displaystyle \genfrac{}{}{0pt}{}{n_a}{b_{a1}}}\right)r_{a1}^{b_{a1}}\left[1r_{a1}\right]^{n_ab_{a1}}\right].`$ (4) Note that the $`n_{D+1}`$ is just a “temporary” variable, which keeps track of the number of $`m_D`$’s who survive the Verhulst “pruning” so that they can give birth before dying from old age. Let us emphasize that this equation is actually quite complex, since $`V`$ is a function of the total population. Multiplying (4) by $`n_c`$ and summing over all the other indices, we obtain $`n_a_{t+1}=V(N)n_{a1}_t,(a>0),`$ (5) $`n_0_{t+1}={\displaystyle \underset{d}{}}F_dr_dV(N)n_d_t,`$ (6) where $`_t`$ denotes the average of $``$ over $`P(𝐧,t)`$. These equations are exact. However, due to the presence of $`N`$ through $`V`$, all moments of $`P`$ may be coupled together. The mean field (MF) approximation consists of replacing the higher order moments by appropriate products of the first moment. Hence we find $`n_a_{t+1}^{\mathrm{MF}}=\left[V({\displaystyle \underset{c}{}}n_c_t^{\mathrm{MF}})\right]n_{a1}_t^{\mathrm{MF}},(a>0),`$ (7) $`n_0_{t+1}^{\mathrm{MF}}=\left[V({\displaystyle \underset{c}{}}n_c_t^{\mathrm{MF}})\right]{\displaystyle \underset{d}{}}F_dr_dn_d_t^{\mathrm{MF}},`$ (8) where, to be clear, we have written the explicit expression for $`N`$. These non-linear equations are known to contain a rich variety of behavior, depending on the details of $`F_c,r_c,`$ and $`V`$. For example, if $`_cF_cr_c<1`$, the reproductive rates are too low and the population will eventually die out. On the other hand, if the reproductive rates are large enough, the population will display period doubling bifurcations and chaos . Let us focus on the “moderate” range, so that a simple non–zero steady–state exists. In that case these equations are easily solved to give $$n_a^{\mathrm{MF}}=N(z)\frac{z^a(1z)}{(1z^{D+1})},$$ (9) where $`z`$ is the unique, positive, real root of the equation $`_cF_cr_cz^{c+1}=1`$, and $`N(z)`$ is the steady state total population, given by the value that satisfies $`V(N)=z`$. Our objective is to go beyond such well–known mean field solutions and investigate fluctuations and correlations, i.e., the second moments of $`P`$. Thus, we multiply Eq. (4) by $`n_an_b`$ and sum over all the other indices. For $`a>0`$, $`b>0`$, we have $`n_an_b_{t+1}=V^2n_{a1}n_{b1}_t+\delta _{ab}V(1V)n_{a1}_t,`$ (10) $`n_an_0_{t+1}={\displaystyle \underset{c}{}}F_cr_cV^2n_{a1}n_c_t+`$ (11) $`+F_{a1}r_{a1}V(1V)n_{a1}_t,`$ (12) $`n_0n_0_{t+1}={\displaystyle \underset{c,d}{}}F_cr_cF_dr_dV^2n_cn_d_t+`$ (13) $`+{\displaystyle \underset{c}{}}\left[F_c^2(r_cVn_c_tr_c^2V^2n_c_t)\right].`$ (14) Note that, like Eqs. (5) and (6), these are exact. Assuming $`N_01`$, and that the system is well away from “critical” points (e.g., bifurcations and the survival/extinction transition), it is reasonable to postulate a Gaussian distribution for $`P^{}\left(𝐧\right)`$ with width of $`O(\sqrt{N_0})`$. Rewriting Eqs. (5,6) and (10-14) for $`^{}`$, we then have a closed set of equations for the first and second moments. This approach should form the first step in a systematic expansion of all quantities in decreasing powers of $`N_0`$. Furthermore, in the same spirit, we will let $`𝐧/N_0`$ assume continuous values. As a check, we will compare the results from this approach with those from a Monte–Carlo simulation, for a simple case. Proceeding, let us write $`P^{}\left(𝐧\right)=\left({\displaystyle \frac{1}{2\pi N_0}}\right)^{(D+1)/2}{\displaystyle \frac{1}{\sqrt{det𝔾}}}\times `$ (15) $`\times \mathrm{exp}\left[{\displaystyle \frac{1}{2N_0}}{\displaystyle \underset{c,d}{}}\left(n_c\overline{n}_c\right)G_{cd}^1\left(n_d\overline{n}_d\right)\right],`$ (16) where we expect the unknown (to be determined) parameters $`\overline{𝐧}`$ and $`𝔾`$ to be of $`O(N_0)`$ and $`O(1)`$, respectively. Note that we will integrate $`𝐧`$ from $`\mathrm{}\mathrm{}`$ rather than from $`0\mathrm{}`$, a simplification which will introduce only negligible errors of $`O(\mathrm{exp}[N_0])`$. Averages can now be computed using $$f(n_a)^{}=f(\overline{n}_a)+\frac{1}{2}\underset{c,d}{}\frac{^2f}{\overline{n}_c\overline{n}_d}N_0G_{cd}+\mathrm{}.$$ (17) Note that the second term in Eq. (17) is expected to be suppressed by a factor of $`O(1/N_0)`$ compared to the first term $`f(\overline{n}_a)`$. Hence the right hand side of Eq. (17) is actually an expansion in powers of $`1/N_0`$. This ordering allows us to set up a systematic perturbation theory, which can be pushed to higher orders if desired. From now on we drop the bars for clarity ($`\overline{n}_an_a`$). Defining $`\xi _cn_c/N_0\text{ and }V^{}dV/d\xi `$, we have $`V/n_c=V^{}(\xi )/N_0`$ and $`^2V/n_cn_d=V^{\prime \prime }(\xi )/N_0^2`$, independent of $`c`$ or $`d`$. Applying Eq. (17) to Eq. (5) gives $$n_a=Vn_{a1}+V^{}\underset{c}{}G_{ca1}+\frac{n_{a1}}{2N_0}V^{\prime \prime }\underset{c,d}{}G_{cd}+\mathrm{}.$$ (18) An equation for $`n_0`$ can be similarly derived. With the assumptions $`𝐧O(N_0)`$ and $`𝔾O(1)`$, the latter two terms in Eq. (18) represent $`O(1/N_0)`$ corrections to the mean field results, while the remaining (lowest order) pieces make up the mean field equation (7). Writing a perturbative expansion: $`n_a=n_a^{(0)}+n_a^{(1)}+\mathrm{}`$(with $`n_a^{(k)}O(N_0^{1k})`$), we see that $`n_a^{(0)}`$ is given by Eq. (9), while the first order result is $`n_a^{(1)}={\displaystyle \underset{c}{}}\left(\left[𝕀S\right]^1\right)_{ac}U_c,\mathrm{with}`$ (19) $`S_{ab}={\displaystyle \frac{[Vn_{a1}]}{n_b}},U_a={\displaystyle \frac{1}{2}}{\displaystyle \underset{c,d}{}}{\displaystyle \frac{^2[Vn_{a1}]}{n_cn_d}}N_0G_{cd},(a>0),`$ (20) $`S_{0b}={\displaystyle \underset{c}{}}F_cr_c{\displaystyle \frac{[Vn_c]}{n_b}},U_0={\displaystyle \frac{1}{2}}{\displaystyle \underset{c,d,e}{}}F_cr_c{\displaystyle \frac{^2[Vn_c]}{n_dn_e}}N_0G_{de},`$ (21) where $`𝕀`$ is the unit matrix and $`𝕊`$ is the stability matrix associated with the mean field (zeroth order) stationary solution. Note that both $`𝕊`$ and $`U`$ need to be evaluated at zeroth order only. With our assumptions about $`F_a,r_a`$, and $`V(N)`$, the eigenvalues of the stability matrix $`𝕊`$ usually lie within the unit circle, implying that our mean field solution is stable. However, for sufficiently high reproductive rates, perturbations with $`\delta n_an_a`$ (i.e., populations with the same relative age distribution, but with with different total numbers of individuals) can have eigenvalues of less than $`1`$. This is the signal of a period doubling bifurcation, leading to the breakdown of our Gaussian perturbation expansion (see also below). Applying the same analysis to the second moments, we obtain, after some lengthy algebra, $$G_{ab}\underset{c,d}{}S_{ac}S_{bd}G_{cd}=K_{ab},a,b,$$ (22) where $`K_{ab}=\delta _{ab}(1V){\displaystyle \frac{n_a}{N_0}},a,b>0,`$ (23) $`K_{a0}=K_{0a}=F_{a1}r_{a1}(1V){\displaystyle \frac{n_a}{N_0}},a>0,`$ (24) $`K_{00}={\displaystyle \frac{1}{N_0}}{\displaystyle \underset{b}{}}VF_b^2r_bn_b(1Vr_b).`$ (25) Again, all quantities need to be evaluated only at the zeroth order, so that, e.g., $`V`$ is just $`z`$. In compact form this equation can be written as $`𝔾𝕊𝔾𝕊^T=𝕂`$, which may be solved by series $$𝔾=𝕂+𝕊𝕂𝕊^T+𝕊^2𝕂\left(𝕊^T\right)^2+\mathrm{}=\underset{n}{}𝕊^n𝕂\left(𝕊^T\right)^n.$$ (26) Since the eigenvalues and eigenvectors of $`𝕊`$ are known, let us write $`𝕊=𝕄𝔼𝕄^1`$, where $`𝔼`$ is in Jordan form, with the eigenvalues on the diagonal, and $`𝕄`$ is the matrix (with its columns) composed of the corresponding right eigenvectors. Note that $`𝕄`$ is not necessarily orthogonal or unitary. If we define $`\stackrel{~}{𝔾}=𝕄^1𝔾\left(𝕄^T\right)^1`$ and $`\stackrel{~}{𝕂}=𝕄^1𝕂\left(𝕄^1\right)^T`$, then it is straightforward to show that $`\stackrel{~}{𝔾}=_n𝔼^n\stackrel{~}{𝕂}𝔼^n`$. For simplicity, let us focus on the case where $`𝔼`$ is diagonal. Then the sum is easily performed, so that $$\stackrel{~}{G}_{ab}=\stackrel{~}{K}_{ab}/\left(1e_ae_b\right)\text{(no sum)},$$ (27) where the $`\{e\}`$ are the eigenvalues. Since $`𝔾=𝕄\stackrel{~}{𝔾}𝕄^T`$, we can directly obtain the matrix $`𝔾`$ and with it all the information about the Gaussian probability distribution (15). The explicit formula for computing $`𝔾`$ is our principal result. Given a particular form of $`V(N)`$ and reproductive parameters $`r_a,F_a`$, we can compute $`𝔾`$ and find the fluctuations in, as well as the correlations between, the populations of various ages. The result (27) contains a further appealing feature: the signal of bifurcation. From stability analysis, we know that period doubling emerges when the eigenvalue associated with $`\delta n_an_a`$ reaches $`1`$. Examining Eq. (27), we see that it is precisely this feature which signals the breakdown of the Gaussian approximation. Furthermore, in many studies of, e.g., the Penna “bit–string” model , menopause sets in before death, so that $`𝔼`$ is not diagonal. Then the final expression for $`\stackrel{~}{𝔾}`$ will be slightly more complicated, although the above conclusions will remain qualitatively unchanged. To check the above analysis, we study the simplest possible case: a $`2`$ age system (i.e. $`D=1`$), with $`r_a=F_a=1`$. Choosing the exponential form for $`V`$ with $`s_0=1`$ and $`N_0=100`$, mean field theory yields $`n_0^{(0)}=157.4`$ and $`n_1^{(0)}=119.3`$. Performing our analysis, we arrive at the first order corrections to $`n_0`$ and $`n_1`$, the fluctuations in the populations of each age, and the correlation between the populations of the two ages. The results are listed in Table I, alongside those from Monte–Carlo simulations . The agreement is excellent, validating our approach. Note that the corrections $`n_a^{(1)}/n_a^{(0)}`$ are less than 1%, vindicating our assertion that they should be $`O(1/N_0)`$. Up to this point we have been considering models with discrete time steps. However it is perfectly possible, and sometimes more appropriate biologically, to analyze models in continuous time . Let us conclude with some brief remarks about fluctuations in this context. A suitable equation for the mean field population dynamics is $$\frac{n(x,t)}{t}=\frac{n(x,t)}{x}\lambda n(x,t)_0^Dn(x^{},t)𝑑x^{},$$ (28) with boundary conditions for birth at $`x=0`$ and certain death at $`x=D`$. However the birth/death/aging processes giving rise to Eq. (28) can also be written as a ballistic reaction model on a discrete spatial lattice but with continuous time. As is well known , starting from the corresponding microscopic lattice master equation, techniques now exist to map this model onto a field theory in continuous space–time. The ensuing action can be recast as a Langevin equation, with the result being Eq. (28), but with extra multiplicative noise terms. The form of these noise terms would then be completely specified, without any ad–hoc guesses. Unfortunately the field–theoretic action is rather awkward, due to the presence of non–local interactions and non–local, multiplicative noise (from fluctuations in the birth process at $`x=0`$). However, simplifications occur if we are interested only in the simple, non–zero steady state, where expansions about the mean field solution should be adequate. In this case the leading noise terms enter additively, so that a perturbation theory analogous to the above approach can be set up. In conclusion, we have developed analytic techniques for dealing with fluctuation effects in a general class of population models with age structure. The results we have presented also form a first step towards an improved analytic understanding of the “bit–string” models of evolution. Finally, it would be interesting to perform further investigations near the bifurcation point, since interesting collective behavior can be expected in that region. We thank R. Desai, B. Schmittmann, and U.C. Täuber for illuminating discussions, and also R. J. Astalos for sharing simulation results prior to publication. This research has been supported in part by the NSERC of Canada and the US National Science Foundation through the Division of Materials Research.
warning/0001/hep-ph0001259.html
ar5iv
text
# 1 Introduction ## 1 Introduction The fragmentations of the $`\mathrm{\Lambda }`$ hyperon, in particular for polarized $`\mathrm{\Lambda }`$, have received a lot of attention recently, both theoretically and experimentally \[1-23\]. There are several reasons for this. First, the spin structure of the $`\mathrm{\Lambda }`$ is rather simple in the naive quark model, since the spin of the $`\mathrm{\Lambda }`$ is completely carried by the valence strange quark, while the up and down quarks form a system in a spin singlet state and give no contribution to the $`\mathrm{\Lambda }`$ spin. Therefore any observation of polarization of the up and down quarks in a $`\mathrm{\Lambda }`$, which departs from this simple picture, would indicate interesting physics related to a novel hadron spin structure . Second, the polarization of the produced $`\mathrm{\Lambda }`$ can be easily determined experimentally by the reconstruction of its decay products. The self-analyzing property due to the characteristic decay mode $`\mathrm{\Lambda }p\pi ^{}`$, with a large branching ratio of 64%, makes it most suitable for studying the fragmentation of various polarized quarks to $`\mathrm{\Lambda }`$. In addition, as pointed out by Gribov and Lipatov , the fragmentation function $`D_q^h(z)`$, for a quark $`q`$ splitting into a hadron $`h`$ with longitudinal momentum fraction $`z`$, can be related to the quark distribution $`q_h(x)`$, for finding the quark $`q`$ inside the hadron $`h`$ carrying a momentum fraction $`x`$, by the reciprocity relation $$D_q^h(z)q_h(x).$$ (1) $`D_q^h`$ and $`q_h`$ depend also on the energy scale $`Q^2`$ and this relation holds, in principle, in a certain $`Q^2`$ range and in leading order approximation. It is important to recall the earlier stronger relation by Drell, Levy and Yan (DLY) , connecting by analytic continuation the DIS structure functions and the fragmentation functions in $`e^+e^{}`$ collisions. For a recent extensive work on the validity of the DLY relation to $`O`$$`(\alpha _s^2)`$, see Ref. , where the Gribov-Lipatov relation Eq. (1) is also verified to hold in leading order for the space- and time-like splitting functions of QCD. Moreover, although Eq. (1) is only valid at $`x1`$ and $`z1`$, it provides a reasonable guidance for a phenomenological parametrization of the various quark to $`\mathrm{\Lambda }`$ fragmentation functions, since we are still lacking a good understanding of the spin and flavor structure of these fragmentation functions. Therefore, at least, we can get some information on the spin and flavor structure of the $`\mathrm{\Lambda }`$ from its fragmentation functions at large $`z`$. The flavor symmetry $`SU(3)_F`$ in the octet baryons can be also used in order to have a deeper insight on the nucleon spin structure. On the other hand, there have been some recent progress in the measurements of the polarized $`\mathrm{\Lambda }`$ production. The longitudinal $`\mathrm{\Lambda }`$ polarization in $`e^+e^{}`$ annihilation at the Z-pole was observed by several Collaborations at CERN . Very recently, the HERMES Collaboration at DESY reported the result of the longitudinal spin transfer to the $`\mathrm{\Lambda }`$ in polarized positron DIS . Also the E665 Collaboration at FNAL measured the $`\mathrm{\Lambda }`$ and $`\overline{\mathrm{\Lambda }}`$ spin transfers from muon DIS , and they observed very different behaviour for $`\mathrm{\Lambda }`$ and $`\overline{\mathrm{\Lambda }}`$ polarizations, though the precision of the data is still rather poor. Several years ago, Brodsky, Burkardt and Schmidt provided a reasonable description of the polarized quark distributions of the nucleon in a pQCD based model . This model has also been successfully used in order to explain the large single-spin asymmetries found in semi-inclusive pion production in $`pp`$ collisions, while other models have not been able to fit the data . In this paper, we extend this analysis to the semi-inclusive production of $`\mathrm{\Lambda }`$ and $`\overline{\mathrm{\Lambda }}`$ in DIS and we try to understand their spin-dependent features. Especially, a possible sea quark and antiquark asymmetry is tested against the E665 experimental results in the $`\mathrm{\Lambda }`$ and $`\overline{\mathrm{\Lambda }}`$ spin transfers. The paper is organized as follows. In section 2 we will present various formulae to derive the $`z`$-dependence of the spin transfer and polarization of the $`\mathrm{\Lambda }`$ ($`\overline{\mathrm{\Lambda }}`$) in lepton DIS. In section 3 we calculate these spin observables in the pQCD based model, and we find that this model gives a good description of the available $`\mathrm{\Lambda }`$ data. In section 4 we present an analysis of the possible contribution from the sea and we try to outline the different trends of the spin transfer for $`\mathrm{\Lambda }`$ and $`\overline{\mathrm{\Lambda }}`$ in the E665 experiment. We test a possible scenario where the sea quarks in the $`\mathrm{\Lambda }`$ ( or sea antiquarks in the $`\overline{\mathrm{\Lambda }}`$ ) are negatively polarized, but the sea antiquarks in the $`\mathrm{\Lambda }`$ ( or sea quarks in the $`\overline{\mathrm{\Lambda }}`$ ) are positively polarized. Finally, we present some concluding remarks in section 5. ## 2 Spin observables in the $`\mathrm{\Lambda }`$ ($`\overline{\mathrm{\Lambda }}`$) fragmentation There are available data on polarized $`\mathrm{\Lambda }`$ ($`\overline{\mathrm{\Lambda }}`$) fragmentation functions in $`e^+e^{}`$ annihilation at the Z-pole and also in lepton DIS. The $`\mathrm{\Lambda }`$ polarization in the $`e^+e^{}`$ annihilation at the Z-pole was previously analysed , and here we concentrate on the spin transfer for the $`\mathrm{\Lambda }`$ production in lepton DIS. For a longitudinally polarized charged lepton beam and an unpolarized target, the $`\mathrm{\Lambda }`$ polarization along its own momentum axis is given in the quark parton model by $$P_\mathrm{\Lambda }(x,y,z)=P_BD(y)A^\mathrm{\Lambda }(x,z),$$ (2) where $`P_B`$ is the polarization of the charged lepton beam, which is of the order of 0.7 or so . $`D(y)`$, whose explicit expression is $$D(y)=\frac{1(1y)^2}{1+(1y)^2},$$ (3) is commonly referred to as the longitudinal depolarization factor of the virtual photon with respect to the parent lepton, and $$A^\mathrm{\Lambda }(x,z)=\frac{\underset{q}{}e_q^2[q^N(x,Q^2)\mathrm{\Delta }D_q^\mathrm{\Lambda }(z,Q^2)+(q\overline{q})]}{\underset{q}{}e_q^2[q^N(x,Q^2)D_q^\mathrm{\Lambda }(z,Q^2)+(q\overline{q})]},$$ (4) is the longitudinal spin transfer to the $`\mathrm{\Lambda }`$. Here $`y=\nu /E`$ is the fraction of the incident lepton’s energy that is transferred to the hadronic system by the virtual photon. We see that the $`y`$-dependence of $`P_\mathrm{\Lambda }`$ factorizes and it can be reduced to a numerical coefficient when $`D(y)`$ is integrated over a given energy range, corresponding to some experimental cuts. In Eq. (4), $`q^N(x,Q^2)`$ is the quark distribution for the quark $`q`$ in the nucleon, $`D_q^\mathrm{\Lambda }(z,Q^2)`$ is the fragmentation function for $`\mathrm{\Lambda }`$ production from quark $`q`$, $`\mathrm{\Delta }D_q^\mathrm{\Lambda }(z,Q^2)`$ is the corresponding longitudinal spin-dependent fragmentation function, and $`e_q`$ is the quark charge in units of the elementary charge $`e`$. In the quark distribution function, $`x=Q^2/2M\nu `$ is the Bjorken scaling variable, $`q^2=Q^2`$ is the squared four-momentum transfer of the virtual photon, $`M`$ is the proton mass, and in the fragmentation function, $`z=E_\mathrm{\Lambda }/\nu `$ is the energy fraction of the $`\mathrm{\Lambda }`$, with energy $`E_\mathrm{\Lambda }`$. In a region where $`x`$ is large enough, say $`0.2x0.7`$, one can neglect the antiquark contributions in Eq. (4) and probe only the valence quarks of the target nucleon. On the contrary, if $`x`$ is much smaller, one is probing the sea quarks and therefore the antiquarks must be considered as well. For $`\overline{\mathrm{\Lambda }}`$ production the polarization $`P_{\overline{\mathrm{\Lambda }}}`$ has an expression similar to Eq. (2), where the spin transfer $`A^{\overline{\mathrm{\Lambda }}}(x,z)`$ is obtained from Eq. (4) by replacing $`\mathrm{\Lambda }`$ by $`\overline{\mathrm{\Lambda }}`$. The $`\mathrm{\Lambda }`$ and $`\overline{\mathrm{\Lambda }}`$ fragmentation functions are related since we can safely assume matter-antimatter symmetry, i.e. $`D_{q,\overline{q}}^\mathrm{\Lambda }(z)=D_{\overline{q},q}^{\overline{\mathrm{\Lambda }}}(z)`$ and similarly for $`\mathrm{\Delta }D_{q,\overline{q}}^\mathrm{\Lambda }(z)`$. It is useful to consider some kinematics variables in the virtual photon-nucleon center-of-mass (c.m.) frame, because the semi-inclusive $`\mathrm{\Lambda }`$ production in DIS can be also regarded as the inclusive hadronic reaction $`\gamma ^{}p\mathrm{\Lambda }X`$. The experimental results are usually presented as functions of the variables $`x_F`$ or $`z^{}`$ rather than $`z`$, so we need to make a general kinematics analysis of the relation of $`x_F`$ and $`z^{}`$ in terms of $`z`$. From the discussion in the Appendix we know that $`x_Fz`$ and $`z^{}z`$ in the Bjorken limit for $`z0`$, and also we know that the produced $`\mathrm{\Lambda }`$ needs to have the same direction as the virtual photon, both in the nucleon rest frame and in the c.m. frame, i.e., the produced $`\mathrm{\Lambda }`$ is collinear with the virtual photon and it is in the current fragmentation region. Therefore we can compare the $`z`$-dependent predictions of the spin transfers with the data expressed in terms of $`x_F`$ or $`z^{}`$. We now turn our attention to the production of any hadron $`h`$ from neutrino and antineutrino DIS processes. The longitudinal polarizations of $`h`$ in its momentum direction, for $`h`$ in the current fragmentation region can be expressed as, $$P_\nu ^h(x,y,z)=\frac{[d(x)+\varpi s(x)]\mathrm{\Delta }D_u^h(z)(1y)^2\overline{u}(x)[\mathrm{\Delta }D_{\overline{d}}^h(z)+\varpi \mathrm{\Delta }D_{\overline{s}}^h(z)]}{[d(x)+\varpi s(x)]D_u^h(z)+(1y)^2\overline{u}(x)[D_{\overline{d}}^h(z)+\varpi D_{\overline{s}}^h(z)]},$$ (5) $$P_{\overline{\nu }}^h(x,y,z)=\frac{(1y)^2u(x)[\mathrm{\Delta }D_d^h(z)+\varpi \mathrm{\Delta }D_s^h(z)][\overline{d}(x)+\varpi \overline{s}(x)]\mathrm{\Delta }D_{\overline{u}}^h(z)}{(1y)^2u(x)[D_d^h(z)+\varpi D_s^h(z)]+[\overline{d}(x)+\varpi \overline{s}(x)]D_{\overline{u}}^h(z)},$$ (6) where the terms with the factor $`\varpi =\mathrm{sin}^2\theta _c/\mathrm{cos}^2\theta _c`$ ($`\theta _c`$ is the Cabibbo angle) represent Cabibbo suppressed contributions. The explicit formulae for the case where $`h`$ is $`\mathrm{\Lambda }`$ and $`\overline{\mathrm{\Lambda }}`$, including only the Cabibbo favored contributions, have been given in . We have neglected the charm contributions both in the target and in hadron $`h`$. One advantage of neutrino (antineutrino) process is that the scattering of a neutrino beam on a hadronic target provides a source of polarized quarks with specific flavor structure, and this particular property makes the neutrino (antineutrino) process an ideal laboratory to study the flavor-dependence of quark to hadron fragmentation functions, especially in the polarized case . The detailed $`x`$-, $`y`$-, and $`z`$\- dependencies can provide more information concerning the various fragmentation functions. As a special case, the $`y`$-dependence can be simply removed by integrating over the appropriate energy range, and the $`x_F`$\- or $`z^{}`$-dependencies of these polarizations can be obtained using the above formulae. It is interesting to notice that, after consideration of symmetries between different quark to $`\mathrm{\Lambda }`$ and $`\overline{\mathrm{\Lambda }}`$ fragmentation functions , there are only eight independent fragmentation functions $$D_q^\mathrm{\Lambda },D_{\overline{q}}^\mathrm{\Lambda },D_q^\mathrm{\Lambda }+\varpi D_s^\mathrm{\Lambda },D_{\overline{q}}^\mathrm{\Lambda }+\varpi D_{\overline{s}}^\mathrm{\Lambda },$$ (7) and $$\mathrm{\Delta }D_q^\mathrm{\Lambda },\mathrm{\Delta }D_{\overline{q}}^\mathrm{\Lambda },\mathrm{\Delta }D_q^\mathrm{\Lambda }+\varpi \mathrm{\Delta }D_s^\mathrm{\Lambda },\mathrm{\Delta }D_{\overline{q}}^\mathrm{\Lambda }+\varpi \mathrm{\Delta }D_{\overline{s}}^\mathrm{\Lambda },$$ (8) where $`q`$ denotes $`u`$ and $`d`$. Different combinations of unpolarized and polarized $`\mathrm{\Lambda }`$ and $`\overline{\mathrm{\Lambda }}`$ productions in neutrino and antineutrino processes and choices of specific kinematics regions with different $`x`$, $`y`$, and $`z`$ can measure the above fragmentation functions efficiently. From Eqs. (7) and (8) it is possible to extract the various strange quark fragmentation functions $$D_s^\mathrm{\Lambda },D_{\overline{s}}^\mathrm{\Lambda },\mathrm{\Delta }D_s^\mathrm{\Lambda },\mathrm{\Delta }D_{\overline{s}}^\mathrm{\Lambda },$$ (9) provided the accuracy of the data is high enough. This supports the conclusion of that hadron production in neutrino (antineutrino) DIS processes provides an ideal laboratory to study the flavor dependence of the quark fragmentation. Another advantage of the neutrino (antineutrino) processes is that the antiquark to $`\mathrm{\Lambda }`$ fragmentation can also be conveniently extracted, and this can be compared to specific predictions concerning the antiquark polarizations inside baryons, which in turn are related to the proton spin problem . To our knowledge, good precision data on $`\mathrm{\Lambda }`$ and $`\overline{\mathrm{\Lambda }}`$ production will be available soon from the NOMAD neutrino beam experiment , thus our knowledge of the various quark to $`\mathrm{\Lambda }`$ fragmentation functions will be improved. The theoretical predictions here may provide a practical guidance for experimental analysis, and a better understanding of the physics observations. ## 3 Description of the $`\mathrm{\Lambda }`$ and $`\overline{\mathrm{\Lambda }}`$ spin observables in a pQCD based model In previous works , the quark distributions of the $`\mathrm{\Lambda }`$ and other octet baryons at large $`x`$ have been discussed in the framework of the pQCD based model. In the region $`x1`$, pQCD can give rigorous predictions for the behavior of distribution functions . In particular, it predicts “helicity retention”, which means that the helicity of a valence quark will match that of the parent nucleon. Explicitly, the quark distributions of a hadron $`h`$ have been shown to satisfy the counting rule , $$q_h(x)(1x)^p,$$ (10) where $$p=2n1+2\mathrm{\Delta }S_z.$$ (11) Here $`n`$ is the minimal number of the spectator quarks, and $`\mathrm{\Delta }S_z=|S_z^qS_z^h|=0`$ or $`1`$ for parallel or anti-parallel quark and hadron helicities, respectively . For the $`\mathrm{\Lambda }`$, we have explicit spin distributions for each valence quark, $$u_v^{}(x)=d_v^{}(x)=\frac{1}{x^{\alpha _v}}[A_{u_v}(1x)^3+B_{u_v}(1x)^4],$$ (12) $$u_v^{}(x)=d_v^{}(x)=\frac{1}{x^{\alpha _v}}[C_{u_v}(1x)^5+D_{u_v}(1x)^6],$$ (13) $$s_v^{}(x)=\frac{1}{x^{\alpha _v}}[A_{s_v}(1x)^3+B_{s_v}(1x)^4],$$ (14) $$s_v^{}(x)=\frac{1}{x^{\alpha _v}}[C_{s_v}(1x)^5+D_{s_v}(1x)^6].$$ (15) Here $`\alpha _v1/2`$ is controlled by Regge exchange for nondiffractive valence quarks. The parameters $`A_{u_v}=1.094`$, $`B_{u_v}=0.677`$, $`C_{u_v}=2.707`$, $`D_{u_v}=2.126`$, $`A_{s_v}=2.188`$, $`B_{s_v}=1.415`$, $`C_{s_v}=2.707`$, and $`D_{s_v}=2.713`$ are taken the same values as in previous paper , with the valence quark helicities $`\mathrm{\Delta }s=0.7`$ and $`\mathrm{\Delta }u=0.1`$. These are slightly different from the Burkardt-Jaffe values $`\mathrm{\Delta }s=0.6`$ and $`\mathrm{\Delta }u=0.2`$, to reflect the fact that the sea quarks might contribute partially to the total $`\mathrm{\Delta }s`$ and $`\mathrm{\Delta }u`$. With these set of parameters, we can have a good description of the $`\mathrm{\Lambda }`$ polarization in the $`e^+e^{}`$ annihilation process at the Z-pole . The calculated $`z`$-dependence of the $`\mathrm{\Lambda }`$ and $`\overline{\mathrm{\Lambda }}`$ polarizations in lepton DIS process with $`x`$ integrated over the range \[0.02, 0.4\] are presented, as dotted curves, in Fig. 1 and Fig. 2 respectively. In our numerical calculations, the CTEQ5 parton distributions of the nucleon at $`Q^2=1.3GeV^2`$ are adopted . We find that our prediction of the $`z`$-dependence of the $`\mathrm{\Lambda }`$ polarization in Fig. 1 is consistent with the HERMES experimental data . The data seem to indicate a trend supporting the prediction of positively polarized $`u`$ and $`d`$ quarks inside $`\mathrm{\Lambda }`$ at large $`x`$. Our prediction is also in qualitative agreement with a Monte Carlo simulation based on inputs of the naive quark model and a model with SU(3) symmetry . The $`z`$-dependence of the fragmentation functions in the Monte Carlo simulation has more validity than naive assumptions without $`z`$-dependence , and this supports our $`z`$-dependence predictions of the quark to $`\mathrm{\Lambda }`$ fragmentation functions based on physics arguments . With the same parton distributions, the $`\mathrm{\Lambda }`$ and $`\overline{\mathrm{\Lambda }}`$ polarizations for the neutrino and antineutrino DIS processes are predicted, as shown, as dotted curves, in Figs. 3 \- 5. We have also compared the situations with and without the Cabibbo suppressed contributions, but we found that the modifications due to the Cabibbo suppression are so small that they can be ignored. However, this small modification from the Cabibbo suppressed terms is specific to the pQCD based model and is not true in general for such situations with very small $`D_q^\mathrm{\Lambda }/D_s^\mathrm{\Lambda }`$ and $`\mathrm{\Delta }D_q^\mathrm{\Lambda }/\mathrm{\Delta }D_s^\mathrm{\Lambda }`$. For example, the contributions from the Cabibbo suppressed terms will be dominant in the situation with $`\mathrm{\Delta }D_q^\mathrm{\Lambda }=0`$. We also point out here that our predictions are dramatically different from the previous calculation based on the assumptions $`\mathrm{\Delta }D_q^\mathrm{\Lambda }(z)=C_q^\mathrm{\Lambda }D_q^\mathrm{\Lambda }(z)`$ where $`C_q^\mathrm{\Lambda }`$ is a constant or a slowly varying function of $`z`$. In our case $`C_q^\mathrm{\Lambda }`$ is a function of $`z`$ and it even changes sign for $`q=u,d`$ when $`z`$ varies between 0 and 1 . The behaviour near $`z=1`$, which contrasts with that of Ref. , is directly related to our explicit parametrization of the quark distributions (see Eqs. (12, 13) ). ## 4 Possible sea quark contributions to the $`\mathrm{\Lambda }`$ and $`\overline{\mathrm{\Lambda }}`$ spin observables We now look at the sea quark contributions to the $`\mathrm{\Lambda }`$ polarization in order to find the rough shapes of the sea quark and sea antiquark in the pQCD based model. Strictly speaking, the Gribov-Lipatov relation Eq. (1) has limitations for its application at small $`x`$ , therefore we should consider our method as a search for a reasonable parametrization of quark to $`\mathrm{\Lambda }`$ fragmentation functions, and then check the validity and relevance of the parametrization by comparing predictions with experimental measurements of various processes. Also there is still a large freedom in the detailed treatments and many assumptions are needed, so that the predictive power at small $`z`$ has more limitations than that at large $`z`$ in our analysis. However, the method has been supported by comparison of the prediction with the experimental data of $`\mathrm{\Lambda }`$ polarizations from both $`e^+e^{}`$-annihilation at the $`Z`$-pole and polarized lepton on the nucleon target DIS scattering , and the uncertainties can be gradually constrained and reduced with more data later on. The sea quark helicity distributions in the pQCD analysis satisfy $$q_s^{}=\frac{1}{x^{\alpha _s}}[A_{q_s}(1x)^5+B_{q_s}(1x)^6],$$ (16) $$q_s^{}=\frac{1}{x^{\alpha _s}}[C_{q_s}(1x)^7+D_{q_s}(1x)^8],$$ (17) $$\overline{q}^{}=\frac{1}{x^{\alpha _s}}[A_{\overline{q}}(1x)^5+B_{\overline{q}}(1x)^6],$$ (18) $$\overline{q}^{}=\frac{1}{x^{\alpha _s}}[C_{\overline{q}}(1x)^7+D_{\overline{q}}(1x)^8],$$ (19) where $`\alpha _s`$, which is controlled by Regge exchange for sea quarks, is taken as the same as that in Ref., i.e. $`\alpha _s1.12`$. We also take the same $`q_s^,`$ and $`\overline{q}^,`$ for $`q=u,d,s`$, for the sake of simplicity. We constrain the sea quark distributions by three conditions: i) $`A_q+B_q=C_q+D_q`$ for $`q=q_s`$ and $`\overline{q}`$ from the convergence of sum rules ; ii) the values of $`\mathrm{\Delta }q_s`$ and $`\mathrm{\Delta }\overline{q}`$; iii) the momentum fractions carried by sea quarks $`x_{q_s}`$ and sea antiquarks $`x_{\overline{q}}`$. This leaves us with one unknown parameter for each set of $`A_q`$, $`B_q`$, $`C_q`$, and $`D_q`$. In the following discussions, the values of $`\mathrm{\Delta }q_s`$ and $`\mathrm{\Delta }\overline{q}`$ are taken so as to let $`\mathrm{\Delta }U=\mathrm{\Delta }u_v+\mathrm{\Delta }q_s+\mathrm{\Delta }\overline{q}=0.2`$ and $`\mathrm{\Delta }S=\mathrm{\Delta }s_v+\mathrm{\Delta }q_s+\mathrm{\Delta }\overline{q}=0.6`$ be consistent with the values of Burkardt-Jaffe sum rule results . It has been pointed out in Ref. that there might be quark-antiquark asymmetry in the quark-antiquark pairs of the nucleon sea, and a possibility to check the strange quark-antiquark asymmetry of the nucleon sea through strange quark and antiquark fragmentations to proton has been also suggested in Ref. . Therefore we can consider the possibility of sea quark-antiquark asymmetry in the $`\mathrm{\Lambda }`$. For comparison we introduce two scenarios, with asymmetric quark-antiquark sea as scenario I, and symmetric quark-antiquark sea as scenario II. For scenario I we choose asymmetric quark-antiquark helicity sums $`\mathrm{\Delta }q_s=0.3`$ and $`\mathrm{\Delta }\overline{q}=0.2`$ with $`x_{q_s}=x_{\overline{q}}0.03`$. We choose $`C_{q_s}`$ as the free parameter and other three parameters are given by the three constrains mentioned before as the solution of $$\begin{array}{ccccc}A_{q_s}=0.988C_{q_s}1.692& & & & \\ B_{q_s}=1.117C_{q_s}+1.915& & & & \\ D_{q_s}=1.129C_{q_s}+0.222& & & & \end{array}$$ (20) The probabilistic interpretation of parton distributions $`q_s^{}`$ and $`q_s^{}`$ implies the rather stringent bounds $$1.714<C_{q_s}<1.718.$$ (21) Similarly, the parameters for the sea antiquarks are constrained by $$\begin{array}{ccccc}A_{\overline{q}}=0.988C_{\overline{q}}+0.860& & & & \\ B_{\overline{q}}=1.117C_{\overline{q}}0.846& & & & \\ D_{\overline{q}}=1.129C_{\overline{q}}+0.0143& & & & \end{array}$$ (22) with $$0<C_{\overline{q}}<0.111.$$ (23) In the following calculation, we take $`C_{q_s}=1.715`$ and $`C_{\overline{q}}=0.1`$. From the solid curves in Figs. 1 and 2, we find that the pQCD based model with sea contributions can reproduce the different behaviors of the $`\mathrm{\Lambda }`$ and $`\overline{\mathrm{\Lambda }}`$ spin transfers, as observed in the E665 experimental data . This implies that the different behaviors with quark-antiquark asymmetry of sea quark fragmentations might be a source for the $`\mathrm{\Lambda }`$ and $`\overline{\mathrm{\Lambda }}`$ spin transfers difference observed by the E665 collaboration. Needless to mention that, although the E665 data is of poor precision, the different behaviors of the $`\mathrm{\Lambda }`$ and $`\overline{\mathrm{\Lambda }}`$ spin transfers might still be a genuine effect. The magnitude of the measured spin transfer Eq. (4) should be less than unity and also $`x_Fz`$ is a good approximation in the kinematics range of the E665 experiment. Also Fig. 3 indicates that there is a big modification to the neutrino induced $`\overline{\mathrm{\Lambda }}`$ polarization in the small and medium $`z`$ region with the sea contributions included, thus $`\mathrm{\Lambda }`$ and $`\overline{\mathrm{\Lambda }}`$ production in neutrino (antineutrino) DIS processes may provide relevant information concerning the antiquark to $`\mathrm{\Lambda }`$ fragmentation functions. However, as has been discussed in , the antiquarks inside the baryons are likely to be unpolarized or slightly positive polarized from a baryon-meson fluctuation model of intrinsic sea quark-antiquark pairs. Therefore we expect that the $`\mathrm{\Lambda }`$ and $`\overline{\mathrm{\Lambda }}`$ productions from neutrino (antineutrino) processes will provide more information about the antiquark polarizations inside baryons, or more precisely, the antiquark to $`\mathrm{\Lambda }`$ fragmentation functions. To reflect the role played by the quark-antiquark asymmetry in reproducing the different behaviors for the fragmentations of $`\mathrm{\Lambda }`$ and $`\overline{\mathrm{\Lambda }}`$ in electron or positron (muon) DIS processes, we introduce scenario II of symmetric quark-antiquark helicity sums $`\mathrm{\Delta }q_s=\mathrm{\Delta }\overline{q}=0.05`$ with $`x_{q_s}=x_{\overline{q}}0.03`$. We have $$\begin{array}{ccccc}A_{q_s}=A_{\overline{q}}=0.988C_q0.416& & & & \\ B_{q_s}=B_{\overline{q}}=1.117C_q+0.534& & & & \\ D_{q_s}=D_{\overline{q}}=1.129C_q+0.118& & & & \end{array}$$ (24) with $$0.422<C_q<0.914.$$ (25) The calculated results with $`C_q=0.6`$ (for both $`q=q_s`$ and $`\overline{q}`$) are also presented in Figs. 1 and 2. From Figs. 1 and 2 we find that in the scenario II we can only produce a small difference of $`\mathrm{\Lambda }`$ and $`\overline{\mathrm{\Lambda }}`$ fragmentations in electron or positron (muon) DIS process. This shows the necessity of having an asymmetric quark-antiquark to $`\mathrm{\Lambda }`$ fragmentations as a possibility to understand the different behaviours of the $`\mathrm{\Lambda }`$ and the $`\overline{\mathrm{\Lambda }}`$ spin transfers in the E665 experiment . We also present in Figs. 3-5 the results from scenario II and notice the different predictions which can be tested in neutrino (antineutrino) DIS processes. ## 5 Concluding remarks In summary, we investigated the $`\mathrm{\Lambda }`$ and $`\overline{\mathrm{\Lambda }}`$ polarizations in lepton DIS in the pQCD based model. We find that the model can give a good description of the available data for the spin transfer. The $`\mathrm{\Lambda }`$ and $`\overline{\mathrm{\Lambda }}`$ polarizations in the neutrino DIS process are predicted. We find that sea contribution gives a big modification to the spin transfer in the small $`z`$ region. The E665 experimental data show very different behaviors of the $`\mathrm{\Lambda }`$ and $`\overline{\mathrm{\Lambda }}`$ spin transfers , which suggests a possible situation such that the sea quarks in the $`\mathrm{\Lambda }`$ ( or sea antiquarks in the $`\overline{\mathrm{\Lambda }}`$) are large negatively polarized, but the sea antiquarks in the $`\mathrm{\Lambda }`$ ( or sea quarks in the $`\overline{\mathrm{\Lambda }}`$) are positively polarized. Acknowledgments: We are very grateful to A. Kotzinian for his useful discussion on the relation between $`z`$ and $`x_F`$. This work is partially supported by Fondecyt (Chile) postdoctoral fellowship 3990048, by the cooperation programmes Ecos-Conicyt and CNRS- Conicyt between France and Chile, by Fondecyt (Chile) grant 1990806 and by a Cátedra Presidencial (Chile), and by National Natural Science Foundation of China under Grant Numbers 19605006, 19875024, 19775051, and 19975052. ## Appendix: a general kinematics analysis We discuss here the kinematics of the semi-inclusive hadron production process from a virtual boson scattering on a hadronic target. For the physics consideration, we take the hadron as $`\mathrm{\Lambda }`$, the boson as photon, and the hadronic target as nucleon, but the discussion applies also to the case of any other hadron, boson, and target. For the general validity of the analysis, we consider the production of the outgoing hadron and the hadronic debris X, treated as an effective particle, regardless of the related sub-process in terms of quarks and gluons. We first consider the kinematics of the particles in the target rest frame. The four-momenta of the incident virtual photon and the nucleon target are $$q=(\nu ,𝐪),p=(M,\mathrm{𝟎}),\mathrm{with}Q^2=𝐪^2\nu ^2,$$ (A.1) and those of the outgoing $`\mathrm{\Lambda }`$ and the hadronic debris with effective mass $`M_X`$ and momentum $`𝐩_X`$ are $$p_\mathrm{\Lambda }=(E_\mathrm{\Lambda },𝐩_\mathrm{\Lambda }),p_X=(E_X,𝐩_X)\mathrm{with}𝐩_X=𝐪𝐩_\mathrm{\Lambda }.$$ (A.2) Using the overall energy conservation we get $$M_X^2=(\nu +ME_\mathrm{\Lambda })^2𝐩_X^2.$$ (A.3) When expressed in terms of $`x=Q^2/2M\nu `$, $`z=E_\mathrm{\Lambda }/\nu `$, and $`\nu `$, we obtain $$M_X^2=2M\nu (1xz)2z\nu ^2\left[1\sqrt{1+\frac{2Mx}{\nu }}\sqrt{1\frac{M_\mathrm{\Lambda }^2}{z^2\nu ^2}}\mathrm{cos}\theta \right]+M^2+M_\mathrm{\Lambda }^2,$$ (A.4) where $`\theta `$ is the angle of $`𝐩_\mathrm{\Lambda }`$ relative to $`𝐪`$. We now consider the kinematics for the particles in the virtual photon and target nucleon c.m. frame where we have $$q^{}=(\nu ^{},𝐪^{}),p^{}=(E^{},𝐪^{}),p_\mathrm{\Lambda }^{}=(E_\mathrm{\Lambda }^{},𝐩_\mathrm{\Lambda }^{}),p_X^{}=(E_X^{},𝐩_\mathrm{\Lambda }^{}).$$ (A.5) We recall that the total energy square of the “two-body” reaction $`\gamma ^{}p\mathrm{\Lambda }X`$ is $$W^2=(p+q)^2=Q^2+2M\nu +M^2=2M\nu (1x)+M^2.$$ (A.6) Standard two-body kinematics gives $$E^{}=\frac{W^2+M^2+Q^2}{2W}=\frac{M\nu +M^2}{W}$$ (A.7) and $$E_\mathrm{\Lambda }^{}=\frac{W^2+M_\mathrm{\Lambda }^2M_X^2}{2W}.$$ (A.8) Now from $$pp_\mathrm{\Lambda }=ME_\mathrm{\Lambda }=p^{}p_\mathrm{\Lambda }^{}=E^{}E_\mathrm{\Lambda }^{}+𝐪^{}𝐩_\mathrm{\Lambda }^{},$$ (A.9) we have $$\mathrm{cos}\theta _{CM}=\frac{ME_\mathrm{\Lambda }E^{}E_\mathrm{\Lambda }^{}}{|𝐪^{}||𝐩_{}^{}{}_{\mathrm{\Lambda }}{}^{}|},$$ (A.10) where $`\theta _{CM}`$ is the angle of $`𝐩_{}^{}{}_{\mathrm{\Lambda }}{}^{}`$ relative to $`𝐪^{}`$. As we have mentioned in section 2, in the experimental measurements one usually introduces two variables attached to the produced particle $`\mathrm{\Lambda }`$: the first one is the Feynman variable $`x_F=2p_L^{}/W`$, where $`p_L^{}=|𝐩_{}^{}{}_{\mathrm{\Lambda }}{}^{}|\mathrm{cos}\theta _{CM}`$ and the second one is the variable $`z^{}`$ defined as $$z^{}=\frac{E_\mathrm{\Lambda }^{}}{E^{}(1x)}.$$ (A.11) From the above kinematics analysis, we can express $`x_F`$ and $`z^{}`$ in terms of $`z`$, $`x`$, $`\nu `$, and $`\theta `$. In the Bjorken limit $`\nu \mathrm{}`$ with $`0<x<1`$, we have from (A.4), for $`z0`$ and $`\theta 0`$, $$M_X^22z\nu ^2(1\mathrm{cos}\theta ),$$ (A.12) which leads to a negative energy $`E_X^{}z\nu ^2(1\mathrm{cos}\theta )/W`$ and is unphysical for $`\theta 0`$. Therefore we must have $`\theta =0`$ and in this case, the Bjorken limit leads to $$M_X^22M\nu (1x)(1z),$$ (A.13) and from (A.7) and (A.8) we have $$|𝐪^{}|E^{}M\nu /W\mathrm{and}|𝐩_{}^{}{}_{\mathrm{\Lambda }}{}^{}|E_\mathrm{\Lambda }^{}M\nu (1x)z/W.$$ (A.14) From Eq. (A.10), we find that $$\mathrm{cos}\theta _{CM}1,$$ (A.15) which implies that the produced $`\mathrm{\Lambda }`$ (or $`\overline{\mathrm{\Lambda }}`$) is also along the virtual photon direction (i.e., $`\theta _{CM}=0`$) in the the c.m. frame. This corresponds to the current fragmentation region since $`x_F>0`$. Finally from the definitions of $`x_F`$ and $`z`$, by using (A.4), (A.7) and (A.8), we immediately find that $$x_Fz\mathrm{and}z^{}z.$$ (A.16)
warning/0001/cond-mat0001083.html
ar5iv
text
# Self-Consistent Tensor Product Variational Approximation for 3D Classical Models ## 1 Introduction The density matrix renormalization group (DMRG) has been widely applied to one-dimensional (1D) quantum systems and two-dimensional (2D) classical systems . A frontier in DMRG is to extend its numerical algorithm to higher dimensional systems, chiefly for 2D quantum and 3D classical systems. As far as the finite system algorithm is concerned, decomposition of higher-dimensional clusters to 1D chains proposed by Liang and Pang works efficiently . On the other hand, we have not obtained any satisfactory answer to extend DMRG toward infinite-size systems in higher dimension. Nishino and Okunishi proposed a way of extending DMRG to 3D classical systems, which they call ‘the corner tensor renormalization group (CTTRG)’ , as a 3D generalization of both the transfer matrix DMRG and the corner transfer matrix renormalization group (CTMRG) for 2D classical systems. Two major problems are found in CTTRG when it is applied to the 3D Ising model. One is that the calculated transition temperature $`T_\mathrm{c}`$ is much higher than one of the most reliable $`T_\mathrm{c}`$ obtained by the Monte Carlo (MC) simulations . The other problem is the very slow decay of the the density-matrix eigenvalues , that spoils the numerical efficiency of the block-spin transformation. A way of generalizing DMRG to higher dimensions is to investigate the variational structure of DMRG, where the variational state for the transfer matrix or the Hamiltonian is written in a product of orthogonal matrices . Two types of 2D tensor product states have been considered as higher-dimensional extensions of this matrix product state. One is ‘the interaction round a face (IRF)’ type product states shown in figure (1(a)) , and the other is the vertex type one in figure (1(b)) . For a 2D tensor product state $`V`$, the variational energy and the variational partition function, respectively, is written as $$\lambda =\frac{V|H|V}{V|V}\mathrm{and}\frac{V|T|V}{V|V},$$ (1) where $`H`$ and $`T`$ denotes a Hamiltonian for a 2D quantum system and a transfer matrix for a 3D classical system. Let us call such a variational estimation as ‘the tensor product variational approximation (TPVA)’ in the following. Calculation of $`\lambda `$ have been performed by MC simulation , product wave function renormalization group (PWFRG) or CTMRG . A key point in TPVA is to find a good variational state $`V`$. So far, they assumed a specific form of $`V`$, which contains several variational parameters, and tried to find out the best $`V`$ by way of the parameter sweep. For example, Okunishi and Nishino investigated TPVA for the 3D Ising model, assuming $`V`$ in the form of the Kramers-Wannier (KW) approximation . Their variational state contains two adjustable parameters, and the best $`V`$ is obtained through a two-parameter sweep. Such an intuitive construction of $`V`$ is, however, not always applicable; for example, we don’t know how to extend the KW approximation for the 3D Potts models . How can we obtain the best $`V`$ automatically for higher-dimensional systems? We find an answer for 3D classical systems. In this paper we propose a self-consistent improvement for the tensor product state $`V`$ by way of CTMRG. We choose the 3D Ising model as an example of the 3D classical systems, and formulate our self-consistent method in terms of the Ising model. In the next section, we introduce the simplest 2D tensor product state, and give the formal expression of the variational partition function $`\lambda `$ in eq.(1). We then obtain the self-consistent equation for $`V`$ in §3, considering the variation $`\delta \lambda /\delta V`$. In §4 we propose a numerical algorithm to solve the self-consistent equation. In §5 we check the numerical efficiency and stability of this algorithm when it is applied to the 3D Ising model and the 3-state ($`q=3`$) 3D Potts model. Conclusions are summarized in §6. ## 2 Tensor Product Variational State We briefly review the variational formulation of TPVA that was used for the KW approximation of the 3D Ising model . Let us consider the 3D Ising model on the simple cubic lattice of the size $`2N\times 2N\times \mathrm{}`$ to $`X`$, $`Y`$ and $`Z`$ directions, respectively, where open (or fixed) boundary conditions are assumed for both $`X`$ and $`Y`$ directions. We are interested in the bulk property of this model, and therefore suppose that the system size $`2N`$ is sufficiently large. Suppose that the neighboring Ising spins $`\sigma `$ and $`\sigma _{}^{}`$ have ferromagnetic interaction $`J\sigma \sigma _{}^{}`$. The transfer matrix $`T`$ from a $`2N\times 2N`$ spin layer $$[\sigma ]\left[\begin{array}{ccccc}\sigma _{\mathrm{1\hspace{0.17em}\hspace{0.17em}\hspace{0.17em}1}}& \mathrm{}& \sigma _{1N}& \mathrm{}& \sigma _{\mathrm{1\hspace{0.17em}\hspace{0.17em}\hspace{0.17em}2}N}\\ \mathrm{}& \mathrm{}& \mathrm{}& \mathrm{}& \mathrm{}\\ \sigma _{N\mathrm{\hspace{0.17em}\hspace{0.17em}1}}& \mathrm{}& \sigma _{NN}& \mathrm{}& \sigma _{N\mathrm{\hspace{0.17em}2}N}\\ \mathrm{}& \mathrm{}& \mathrm{}& \mathrm{}& \mathrm{}\\ \sigma _{2N\mathrm{\hspace{0.17em}1}}& \mathrm{}& \sigma _{2NN}& \mathrm{}& \sigma _{2N\mathrm{\hspace{0.17em}2}N}\end{array}\right]$$ (2) to the next layer $`[\overline{\sigma }]`$ is then expressed as a product of local factors $$T[\overline{\sigma }|\sigma ]=\underset{i=1}{\overset{2N1}{}}\underset{j=1}{\overset{2N1}{}}X_{ij}^{}\underset{ij}{}X_{ij}^{},$$ (3) where $`X_{ij}^{}`$ represents the Boltzmann factor for a local cube $`X_{ij}^{}=\mathrm{exp}\{{\displaystyle \frac{K}{4}}`$ $`(`$ $`\overline{\sigma }_{i^{}j}\overline{\sigma }_{ij}+\overline{\sigma }_{ij^{}}\overline{\sigma }_{ij}+\overline{\sigma }_{i^{}j^{}}\overline{\sigma }_{i^{}j}+\overline{\sigma }_{i^{}j^{}}\overline{\sigma }_{ij^{}}.`$ (4) . $`+`$ $`\overline{\sigma }_{ij}\sigma _{ij}^{}+\overline{\sigma }_{i^{}j}\sigma _{i^{}j}^{}+\overline{\sigma }_{ij^{}}\sigma _{ij^{}}^{}+\overline{\sigma }_{i^{}j^{}}\sigma _{i^{}j^{}}^{}.`$ . $`+`$ $`\sigma _{i^{}j}^{}\sigma _{ij}^{}+\sigma _{ij^{}}^{}\sigma _{ij}^{}+\sigma _{i^{}j^{}}^{}\sigma _{i^{}j}^{}+\sigma _{i^{}j^{}}^{}\sigma _{ij^{}}^{})\}`$ parameterized by $`K=J/k_\mathrm{B}^{}T`$. We have used the notation $`i^{}=i+1`$ and $`j^{}=j+1`$. (See figure (2).) We define $`T[\overline{\sigma }|\sigma ]`$ so that it is symmetric, because the symmetry simplifies the following formulation. The variational state in TPVA is a uniform product of local tensors. In this paper, we focus on the simplest construction of the tensor product state $$V[\sigma ]=\underset{ij}{\overset{}{}}W_{ij}^{}=\underset{ij}{}W\left(\begin{array}{cc}\sigma _{ij}^{}& \sigma _{ij^{}}^{}\\ \sigma _{i^{}j}^{}& \sigma _{i^{}j^{}}^{}\end{array}\right),$$ (5) where the local tensor $`W_{ij}^{}`$ does not contains auxiliary variables, which are shown by black squares in figure (1(a)) ; to include the auxiliary variables is straightforward, but makes the following equations rather lengthy. The tensor product state $`V[\sigma ]`$ is uniform in the sense that $`W_{ij}^{}`$ is position independent. The local tensor $`W_{ij}^{}`$ has 16 parameters, but not all of them are physically independent . For the book keeping, let us use the notation $$\{\sigma _{ij}^{}\}\left(\begin{array}{cc}\sigma _{ij}^{}& \sigma _{ij^{}}^{}\\ \sigma _{i^{}j}^{}& \sigma _{i^{}j^{}}^{}\end{array}\right)$$ (6) for the plaquett spin, and write the local tensor $`W_{ij}^{}`$ simply as $`W\{\sigma _{ij}^{}\}`$. In the same manner, let us write $`X_{ij}^{}`$ as $`X\{\overline{\sigma }_{ij}|\sigma _{ij}^{}\}`$. (See figure (2).) Using $`T[\overline{\sigma }|\sigma ]`$ and $`V[\sigma ]`$ thus defined, the variational partition function per layer is expressed as $`\lambda `$ $`=`$ $`{\displaystyle \frac{{\displaystyle \underset{[\overline{\sigma }][\sigma ]}{\overset{}{}}}V[\overline{\sigma }]T[\overline{\sigma }|\sigma ]V[\sigma ]}{{\displaystyle \underset{[\sigma ]}{\overset{}{}}}\left(V[\sigma ]\right)^2}}={\displaystyle \frac{{\displaystyle \underset{[\overline{\sigma }][\sigma ]}{\overset{}{}}}{\displaystyle \underset{ij}{}}W\{\overline{\sigma }_{ij}^{}\}X\{\overline{\sigma }_{ij}|\sigma _{ij}^{}\}W\{\sigma _{ij}^{}\}}{{\displaystyle \underset{[\sigma ]}{\overset{}{}}}{\displaystyle \underset{ij}{}}\left(W\{\sigma _{ij}^{}\}\right)^2}}`$ (7) $`=`$ $`{\displaystyle \frac{{\displaystyle \underset{[\overline{\sigma }][\sigma ]}{\overset{}{}}}{\displaystyle \underset{ij}{}}G_{}^1\{\overline{\sigma }_{ij}|\sigma _{ij}^{}\}}{{\displaystyle \underset{[\sigma ]}{\overset{}{}}}{\displaystyle \underset{ij}{}}G_{}^0\{\sigma _{ij}^{}\}}}{\displaystyle \frac{Z_{}^1}{Z_{}^0}},`$ where we have defined $`G_{}^0`$ and $`G_{}^1`$ as $`G_{}^0\{\sigma _{ij}^{}\}`$ $`=`$ $`\left(W\{\sigma _{ij}^{}\}\right)^2,`$ $`G_{}^1\{\overline{\sigma }_{ij}|\sigma _{ij}^{}\}`$ $`=`$ $`W\{\overline{\sigma }_{ij}^{}\}X\{\overline{\sigma }_{ij}|\sigma _{ij}^{}\}W\{\sigma _{ij}^{}\}.`$ (8) It should be noted that $`Z_{}^0`$ is a partition function of an IRF model on $`2N\times 2N`$ square, and $`Z_{}^1`$ is that of a 2-layer lattice model of the same size. ## 3 Self-Consistent Relation for the variational state Now we explain the self-consistent equation for the variational state $`V[\sigma ]`$, the equation which is satisfied when $`\lambda `$ in eq.(7) is maximized. Let us consider the variation of $`\lambda `$ with respect to the variations of local tensors $$\frac{\delta \lambda }{\delta V}\underset{ij}{\overset{}{}}\frac{\delta \lambda }{\delta W_{ij}^{}}$$ (9) under the condition that the system size $`2N`$ is sufficiently large and the boundary effect is negligible. Then most of the terms in the r.h.s. are almost the same, and it is sufficient to consider the variation of $`\lambda `$ with respect to the local change $`W_{NN}^{}`$ $`W_{NN}^{}+\delta W_{NN}^{}`$ at the center of the system, where $`W_{NN}^{}`$ represents the local tensor at the center. (See eqs.(2) and (5).) The variation $`\delta \lambda /\delta W_{NN}^{}`$ can be explicitly written down by use of two matrices. One is the diagonal matrix $$A\{\sigma _{NN}^{}\}=\underset{[\sigma ]^{}}{\overset{}{}}\underset{(ij)(NN)}{\overset{}{}}G_{}^0\{\sigma _{ij}^{}\},$$ (10) where $`_{[\sigma ]^{}}^{}`$ denotes spin configuration sum for all the spins in the layer $`[\sigma ]`$ except for the spin plaquett $`\{\sigma _{NN}^{}\}`$ at the center; we interpret $`A\{\sigma _{NN}^{}\}`$ as a 16-dimensional matrix $`M\{\overline{\sigma }_{NN}^{}|\sigma _{NN}^{}\}`$ where $`M\{\sigma _{NN}^{}|\sigma _{NN}^{}\}=A\{\sigma _{NN}^{}\}`$ and is zero when $`\{\overline{\sigma }_{NN}^{}\}\{\sigma _{NN}^{}\}`$. From the definition, $`Z_{}^0`$ in eq.(7) is equal to $`_{\{\sigma _{NN}^{}\}}^{}G_{}^0\{\sigma _{NN}^{}\}A\{\sigma _{NN}^{}\}`$. The other matrix is $$B\{\overline{\sigma }_{NN}^{}|\sigma _{NN}^{}\}=X\{\overline{\sigma }_{NN}|\sigma _{NN}^{}\}\underset{[\overline{\sigma }]^{}[\sigma ]^{}}{\overset{}{}}\underset{(ij)(NN)}{\overset{}{}}G_{}^1\{\overline{\sigma }_{ij}|\sigma _{ij}^{}\},$$ (11) which is related to $`Z_{}^1`$ as $`Z_{}^1=_{\{\overline{\sigma }_{NN}^{}\}\{\sigma _{NN}^{}\}}^{}W_{}\{\overline{\sigma }_{NN}^{}\}B\{\overline{\sigma }_{NN}^{}|\sigma _{NN}^{}\}W\{\sigma _{NN}^{}\}`$. By use of $`A\{\sigma _{NN}^{}\}`$ and $`B\{\overline{\sigma }_{NN}^{}|\sigma _{NN}^{}\}`$ thus created, we can write down $`\lambda `$ as $$\lambda =\frac{Z_{}^1}{Z_{}^0}=\frac{{\displaystyle \underset{\{\overline{\sigma }\}\{\sigma \}}{\overset{}{}}}W\{\overline{\sigma }\}B\{\overline{\sigma }|\sigma \}W\{\sigma \}}{{\displaystyle \underset{\{\sigma \}}{\overset{}{}}}W\{\sigma \}A\{\sigma \}W\{\sigma \}},$$ (12) where we have dropped the subscripts from $`\{\sigma _{NN}^{}\}`$ and $`\{\overline{\sigma }_{NN}\}`$ for book keeping. The condition $`\delta \lambda /\delta W\{\sigma _{NN}^{}\}`$ $`=0`$ draws the eigenvalue problem $$\underset{\{\sigma \}}{\overset{}{}}\frac{1}{A\{\overline{\sigma }\}}B\{\overline{\sigma }|\sigma \}W\{\sigma \}=\lambda W\{\overline{\sigma }\}$$ (13) between the matrix $`A_{}^1B`$ and $`W`$; here we regard $`W\{\sigma \}`$ as a 16-dimensional vector. This is the self-consistent equation that an optimized tensor product state $`V[\sigma ]`$ should satisfy. ## 4 Numerical Algorithm of the Self-Consistent TPVA To use the self-consistent relation eq.(13), we have to obtain $`A\{\sigma \}`$ and $`B\{\overline{\sigma }|\sigma \}`$ for very large $`N`$. Though it is impossible to obtain $`A\{\sigma \}`$ and $`B\{\overline{\sigma }|\sigma \}`$ exactly, the CTMRG enables us to numerically obtain them very accurately. Let us introduce a new notation $$\mu _{ij}^{}(\overline{\sigma }_{ij},\sigma _{ij}^{}),$$ (14) which groups a pair of adjacent spins $`\overline{\sigma }_{ij}`$ and $`\sigma _{ij}^{}`$. Using $`\mu _{ij}^{}`$, we can rewrite the stack of two plaquett spins $`\{\overline{\sigma }_{ij}|\sigma _{ij}^{}\}`$ as $$\{\mu _{ij}^{}\}=\left(\begin{array}{cc}\mu _{ij}^{}& \mu _{ij^{}}^{}\\ \mu _{i^{}j}^{}& \mu _{i^{}j^{}}^{}\end{array}\right),$$ (15) $`X\{\overline{\sigma }_{ij}|\sigma _{ij}^{}\}`$ as $`X\{\mu _{ij}\}`$, and $`G_{}^1\{\overline{\sigma }_{ij}|\sigma _{ij}^{}\}`$ as $`G_{}^1\{\mu _{ij}\}`$. (See figure (2)) We drop the subscripts from $`\{\mu _{ij}^{}\}`$ to write it as $`\{\mu \}`$ when its position is apparent. The matrices $`A\{\sigma \}`$ and $`B\{\mu \}=B\{\overline{\sigma }|\sigma \}`$ can be expressed as a combination of the corner transfer matrices (CTMs) and the half-row transfer matrices (HRTMs), that appears when we apply CTMRG to both the denominator and the numerator of eq.(7) to obtain $`Z_{}^0`$ and $`Z_{}^1`$ . Let us write the CTM used for the calculation of $`Z_{}^0`$ and $`Z_{}^1`$, respectively, as $`C_{}^0(\xi \sigma \xi _{}^{})`$ and $`C_{}^1(\zeta \mu \zeta _{}^{})`$, where $`\xi `$, $`\xi _{}^{}`$, $`\zeta `$, and $`\zeta _{}^{}`$ are $`m`$-state block spin variables. Also let us write HRTM as $`P_{}^0(\xi \sigma \sigma _{}^{}\xi _{}^{})`$ and $`P_{}^1(\zeta \mu \mu _{}^{}\zeta _{}^{})`$ in the same manner. Note that $`C_{}^0(\xi \sigma \xi _{}^{})`$ and $`P_{}^0(\xi \sigma \sigma _{}^{}\xi _{}^{})`$ are created from $`G^0\{\sigma \}`$, and $`C_{}^1(\zeta \mu \zeta _{}^{})`$ and $`P_{}^1(\zeta \mu \mu _{}^{}\zeta _{}^{})`$ are from $`G^1\{\mu \}`$. Combining these CTMs and HRTMs, $`A\{\sigma \}`$ and $`B\{\mu \}`$ are constructed as $`A\{\sigma \}={\displaystyle \underset{\xi _1\mathrm{}\xi _8}{\overset{}{}}}`$ $`P_{}^0(\xi _1^{}\sigma _a^{}\sigma _b^{}\xi _2^{})C_{}^0(\xi _2^{}\sigma _b^{}\xi _3^{})P_{}^0(\xi _3^{}\sigma _b^{}\sigma _c^{}\xi _4^{})C_{}^0(\xi _4^{}\sigma _c^{}\xi _5^{})`$ $`P_{}^0(\xi _5^{}\sigma _c^{}\sigma _d^{}\xi _6^{})C_{}^0(\xi _6^{}\sigma _d^{}\xi _7^{})P_{}^0(\xi _7^{}\sigma _d^{}\sigma _a^{}\xi _8^{})C_{}^0(\xi _8^{}\sigma _a^{}\xi _1^{}),`$ $`B\{\mu \}=X\{\mu \}{\displaystyle \underset{\zeta _1\mathrm{}\zeta _8}{\overset{}{}}}`$ $`P_{}^1(\zeta _1^{}\mu _a^{}\mu _b^{}\zeta _2^{})C_{}^1(\zeta _2^{}\mu _b^{}\zeta _3^{})P_{}^1(\zeta _3^{}\mu _b^{}\mu _c^{}\zeta _4^{})C_{}^1(\zeta _4^{}\mu _c^{}\zeta _5^{})`$ (16) $`P_{}^1(\zeta _5^{}\mu _c^{}\mu _d^{}\zeta _6^{})C_{}^1(\zeta _6^{}\mu _d^{}\zeta _7^{})P_{}^1(\zeta _7^{}\mu _d^{}\mu _a^{}\zeta _8^{})C_{}^1(\zeta _8^{}\mu _a^{}\zeta _1^{})`$ where the positions of the spin variables are shown in figure (3). In principle, we can use $`A\{\sigma \}`$ and $`B\{\mu \}`$ thus constructed to solve the self-consistent eq.(13). To make the self-consistent improvement for $`W\{\sigma \}`$ more efficiently, we employ a numerical algorithm that simultaneously performs the extension of the system size in CTMRG and the self-consistent improvement by eq.(13). The numerical procedures are as follows: * Create $`G_{}^0\{\sigma \}`$ and $`G_{}^1\{\mu \}`$ from $`X\{\mu \}`$ defined in eq.(2) and the initial $`W\{\sigma \}`$: $`G_{}^0\{\sigma \}`$ $`=`$ $`\left(W\{\sigma \}\right)^2,`$ (17) $`G_{}^1\{\mu \}G_{}^1\{\overline{\sigma }|\sigma \}`$ $`=`$ $`W\{\overline{\sigma }\}X\{\overline{\sigma }|\sigma \}W\{\sigma \}.`$ The choice of the initial $`W\{\sigma \}`$ is not so relevant, since it is improved afterward. * Create the initial $`C_{}^0(\xi \sigma \xi _{}^{})`$ and the initial $`P_{}^0(\xi \sigma \sigma _{}^{}\xi _{}^{})`$ from $`G_{}^0\{\sigma \}`$, following the standard initialization procedure in CTMRG . Also create $`C_{}^1(\zeta \mu \zeta _{}^{})`$ and $`P_{}^1(\zeta \mu \mu _{}^{}\zeta _{}^{})`$ from $`G_{}^1\{\mu \}`$ in the same way. * Obtain the matrices $`A\{\sigma \}`$ and $`B\{\mu \}=B\{\overline{\sigma }|\sigma \}`$ using eqs.(16). * Improve $`W\{\sigma \}`$ by multiplying $`A_{}^1B`$ $$W_{\mathrm{new}}^{}\{\overline{\sigma }\}=\underset{\{\sigma \}}{\overset{}{}}\frac{1}{A\{\overline{\sigma }\}}B\{\overline{\sigma }|\sigma \}W_{\mathrm{old}}^{}\{\sigma \},$$ (18) and normalize $`W_{\mathrm{new}}^{}\{\sigma \}`$ so that $$\underset{\{\sigma \}}{\overset{}{}}\left(W_{\mathrm{new}}^{}\{\sigma \}\right)^2=1$$ (19) is satisfied. * Recreate $`G_{}^0\{\sigma \}`$ and $`G_{}^1\{\mu \}`$ by substituting $`W_{\mathrm{new}}^{}\{\sigma \}`$ into eqs.(17). * Extend $`P_{}^0`$ and $`P_{}^1`$ to obtain $`P_{\mathrm{ext}}^0`$ and $`P_{\mathrm{ext}}^1`$, respectively, by joining the recreated $`G_{}^0`$ and $`G_{}^1`$ as shown in figure (4); the numerical details are shown in ref. . Also extend $`C_{}^0`$ and $`C_{}^1`$ to obtain $`C_{\mathrm{ext}}^0`$ and $`C_{\mathrm{ext}}^1`$. * Create density matrices from the extended CTMs, and diagonalizing them to obtain the RG transformations $`\xi _{\mathrm{old}}^{}\sigma \xi _{\mathrm{new}}^{}`$ and $`\zeta _{\mathrm{old}}^{}\mu \zeta _{\mathrm{new}}^{}`$, where $`\xi `$ and $`\zeta `$ are $`m`$-state block spins. Then apply the RG transformations to $`P_{\mathrm{ext}}^0`$, $`P_{\mathrm{ext}}^1`$ $`C_{\mathrm{ext}}^0`$ and $`C_{\mathrm{ext}}^1`$. * Goto (c), and repeat (c)-(g) to improve $`W\{\sigma \}`$ iteratively, and stop when $`W\{\sigma \}`$ reaches its fixed point. To summarize, we put three additional steps (a), (c) and (d) to the standard CTMRG algorithm. ## 5 Numerical Results Let us check the numerical efficiency and stability of the self-consistent TPVA through trial applications to the 3D Ising model and the ferromagnetic $`q=3`$ Potts model. Figure (5) shows the spontaneous magnetization $`\sigma `$ at the center of the $`2N\times 2N\times \mathrm{}`$ system, where the curve, cross marks, and triangles, respectively, represent the result of the MC simulation by Tarpov and Blöte , KW approximation , and the self-consistent TPVA. We calculate $`\sigma `$ after repeating the iteration (c)-(g) in the last section for $`N=10000`$ times at most, keeping $`m=10`$ to $`m=20`$ states for the block spin variables; the convergence with respect to $`m`$ is very fast, where we obtain almost the same $`\sigma `$ for the cases $`m=10`$ and $`20`$. The self-consistent improvement by eq.(18) is monotonous in the whole parameter range, and no oscillatory instability is observed. The calculated transition point $`K_\mathrm{c}=0.2188`$ is about 1.3% smaller than the MC result $`K_\mathrm{c}^{\mathrm{MC}}=0.2216544`$. It turns out that the spontaneous magnetization calculated by KW approximation, which gives the transition point $`K_\mathrm{c}^{\mathrm{KW}}=0.2180`$, is quite close to the result of the self-consistent (SC) TPVA. This means that the intuitive choice of the variational state in KW approximation is actually very good, within the simplest product state defined in eq.(5). Inclusion of auxiliary variables to the tensor product state is necessary for the further improvement of the tensor product variational state. Note that the computational time required for the KW approximation is several times larger than the self-consistent TPVA, because the former finds the partition function extremum via 2-parameter sweep. Figure (6) shows the energy per bond $`E=\delta (\sigma ,\sigma _{}^{})`$ of the ferromagnetic 3D $`q=3`$ Potts model, which is calculated by TPVA keeping $`m`$ up to 15. The self-consistent improvement by eq.(18) is again monotonous, and $`N=1000`$ is sufficient to get the converged data; we need smaller $`N`$ for the Potts model than Ising model, because the phase transition of the Potts model is first order. The calculated energy per bond jumps from $`E^+=0.5173`$ to $`E^{}=0.5933`$ at the calculated transition point $`K_\mathrm{c}=0.54956`$, where the calculated free energy of the disordered phase coincides with that of the ordered phase. The calculated transition point is about 0.18% smaller than one of the most reliable MC result $`K_\mathrm{c}^{\mathrm{MC}}=0.550565\pm 0.000010`$ . The latent heat $`l=3(E^+E^{})=0.22769`$ is about 41% larger than the MC result $`l=0.16160\pm 0.00047`$ . ## 6 Conclusion We have proposed a self-consistent TPVA, which gives the optimized tensor product state for 3D classical systems, by way of the self-consistent improvement of the local tensors. Since the method finds out the best variational state without using a priori knowledge of the system, the self-consistent TPVA is applicable for various 3D models described by short range interactions. To generalize the self-consistent TPVA to 2D quantum systems is a next subject that one might consider. This generalization is not trivial, since we have used the specific property of 3D classical systems when we obtain the self-consistent equation. T. N. thank to G. Sierra and M.A. Martín-Delgado for the discussion about the tensor product state at CSIC. K. O. is supported by JSPS Research Fellowships for Young Scientists. This work was partially supported by the “Research for the Future” Program from The Japan Society for the Promotion of Science (JSPS-RFTF97P00201) and by the Grant-in-Aid for Scientific Research from Ministry of Education, Science, Sports and Culture (No. 09640462 and No. 11640376). Most of the numerical calculations were done by Compaq Fortran on the HPC Alpha21264 Linux workstation.
warning/0001/cond-mat0001117.html
ar5iv
text
# 1 Introduction ## 1 Introduction One of the customary ways in dealing with the volatility smile and term-structure as observed, for example, in the implied Black-Scholes volatilities in the options markets, is to translate the different Black-Scholes volatilities into a ‘local’ volatility (‘implied tree’) . This assumes the stock-price follows the process $$\frac{dS}{S}=\mu (t)dt+\sigma (S(t),t)dW,$$ where $`\sigma `$ is a deterministic function of the stock-price and time, calibrated such that the model matches the observed vanilla options prices, and $`W`$ is a standard Brownian motion (SBM) and the only random source of the model. While the earliest incarnations of this idea led to very erratic behaviour of the local volatility surface, much recent work has focused on how to compute a surface that is both reasonably smooth and fast to find . Both the names DVF (deterministic volatility function) and LVF (local volatility function) are customary for this approach. We will use the abbreviation DLVF, for obvious reasons, and to distinguish it from SLVF (stochastic local volatility function, or stochastic implied tree models), which we shall also discuss in this paper. We shall use LVF when we refer to both classes of models. DLVF models are the preferred skew models of a number of institutions, given that alternative approaches to create a skew effect encounter limitations as to how well they can fit market data . The commonly perceived downside of DLVF models are practical limitations in measuring the volatility exposure of an exotic options portfolio. There are, however, no theoretical grounds on which to affirm that volatility risk can be hedged at all in this framework: A truly deterministic volatility function must be assumed. The two most pressing theoretical reservations one may harbour towards local volatility models, and the problems arising from these concerns, are therefore the following. * When the stock-price moves, but the LVF is assumed to be constant, this imposes a rule on the Black-Scholes-implied volatilities which is neither of the two popular rules nick-named ‘sticky strike’ and ‘sticky delta’. The ‘sticky-implied-tree’ rule leads to different stock-deltas than either of the other two rules . Therefore, before putting trust in local volatility models, we must either justify the emergence of this rule, or show that the validity or otherwise of the ‘sticky-implied-tree’ assumption is immaterial to the P&L of a hedged strategy, which can only be the case if we can hedge against stock and volatility movements separately. * In general, it is difficult to estimate how the local volatility surface moves when a limited number of Black-Scholes-implied volatilities change in the market. Note, however, that measuring this exposure is one of the primary motivations for using a skew model at all. One of the key arguments used in defence of DLVF models is that volatility risk is already priced into the market, and that the DLVF model is only an effective theory which makes use of the fact that volatility risk is already integrated out in the right way by the market-implied local DLVF. This is true for a well-diversified portfolio of vanilla options, but the implied volatility does not compensate sufficiently for option-specific vega risk of vanilla or exotic options. A $`\mathrm{\Delta }`$-hedged portfolio short a log-contract has a much more predictable P&L in the real world than a $`\mathrm{\Delta }`$-hedged portfolio short just one at-the-money vanilla option. This is because in the first case the unknown P&L component is only a function of the realised asset price variance (assuming a diffusive asset price process), whereas in the second case the P&L is also very much a function of the average $`\mathrm{\Gamma }`$ experienced over the life of the option. So, although implied volatilities often compensate the option seller for the vega-risk taken, they do so insufficiently for someone who does not diversify his volatility risk across various strikes. If we want to sell an exotic option, and hope that it is priced in line with the market, we must also show that we either can diversify our vega exposure in a similar way, and that the vega-risk which can not be diversified away is priced in line with the vega-risk premium that is already integrated out in the observable Black-Scholes implied volatilities. Obviously, the DLVF model does not give us a volatility risk premium above the premium implied in the vanilla prices. This is what should be meant when it is claimed that the implied tree values exotics ‘in line with the market’. The concern that needs to be addressed is thus whether the expectations that are already integrated out in the ‘vanilla-implieds’ can be re-used as ‘exotic-implieds’. Note that the only way we can be certain that the effective integration of volatility risk can be applied to exotics is by showing that volatility risk of an exotic option can be hedged with a dynamic trading strategy of vanilla options. Unless one can show that volatility exposure can be hedged against any of the points on the local volatility surface (or any of the variables which parameterise this surface), we may not be entitled to use risk-neutral valuation techniques in a world where implied volatilities are subject to change. This would make the DLVF models unviable. We will demonstrate in this article that these two concerns can be redressed in an arbitrage-free market only if the future spot-volatilities are indeed deterministically identical to their DLVF values. That they are not is clear from empirical work and from simple intuition . The structure of the paper is as follows: First, we give a general discussion of what conditions we might expect in order for the local volatility model to be viable under risk-neutral valuation with an only mildly time-varying volatility function. We observe that these requirements destroy the no-arbitrage condition, and that we thus need to consider the full SLVF model known also as the stochastic implied tree. This paper can be described as a practical guide to (a) the problem of incompleteness and (b) the HJM-like no-arbitrage condition of stochastic implied trees as described in . The above-mentioned early sections of the paper introduce the theoretical problems. Although the way of presenting these problems may be new, the incompleteness problem has been discussed in much greater detail and mathematical rigor in the literature on stochastic volatility models. After this practitioner’s guide to the theoretical problems, we present a discrete framework in which we are able to construct, if it exists, the replicating trading strategy of any exotic option, involving stock and Arrow-Debreu option positions. We demonstrate that this portfolio always exists in a two-step trinomial world. Next, we analyse a three-step trinomial world and find that the stochastic process required to make the hedge work violates the no-arbitrage condition. This problem persists in the continuum limit. We then present the conclusion that in a DLVF model risk-neutral valuation of exotics (without additional arbitrary parameters like for example option-dependent volatility risk premia, which can not be calibrated to fit the vanilla options market) is inconsistent with the assumptions of no arbitrage and unknown future volatility. ## 2 Stochastic Processes with and without Quadratic Variation Some of the issues discussed here benefit from a brief tour through familiar territory. The Black-Scholes world is based on the assumption that the stock price $`S`$ follows the process $$dS=rSdt+\sigma SdW,$$ where the $`dW`$ are the increments of an SBM. The Black-Scholes PDE $$\frac{f}{t}+rS\frac{f}{S}+\frac{1}{2}\sigma ^2S^2\frac{^2f}{S^2}=rf,$$ in a slightly different notation, is called the continuous trading equation $$\mathrm{\Theta }dt+rS\mathrm{\Delta }dt+\frac{1}{2}\sigma ^2S^2\mathrm{\Gamma }dt=frdt.$$ The left-hand side of the continuous trading equation is equal to $`df\mathrm{\Delta }dS+r\mathrm{\Delta }Sdt`$ , the P&L of a $`\mathrm{\Delta }`$-hedged portfolio, which is risk-less in the Black-Scholes world, because the volatility $`\sigma `$ is a known constant. For the rest of the discussion, it will be easier to think of the continuous trading equation in terms of a futures-hedge, as the funding term then drops out, i.e. we think of the continuous trading equation as $$df\frac{f}{F}dF=rfdt,$$ or, after expanding the total differential for $`df(F,t)`$ to first order in $`dt`$, $$\mathrm{\Theta }dt+\frac{1}{2}\sigma ^2F^2\mathrm{\Gamma }_Fdt=rfdt.$$ The continuous trading equation thus tells us that if all the other assumptions of the Black-Scholes world hold, the instantaneous P&L of a long, $`\mathrm{\Delta }`$-hedged, option position is a trade-off between the loss of $`\mathrm{\Theta }`$ on the option and the gain on $`\mathrm{\Gamma }`$, which is proportional to the instantaneous variance rate $`\sigma ^2`$ of the stock-price. The fact that we have a $`\mathrm{\Gamma }`$-related P&L component is a consequence of the fact that the stock-price has non-vanishing quadratic variation $`dS,dF=𝒪(\sqrt{dt})`$. Now imagine that $`S`$ is stochastic, but of vanishing quadratic variation, for example $`dS\left(r+W\frac{t^2}{2}\right)Sdt`$ or $`dF\left(W\frac{t^2}{2}\right)Fdt`$, then the left-hand side of the continuous trading equation is $`df\mathrm{\Delta }_FdF`$ or $$\mathrm{\Theta }dt=rfdt,$$ so the $`\mathrm{\Gamma }`$-related P&L disappears, even though the asset price process is stochastic. The asset price process has a fundamental problem, however: Even though initially it has the desirable property of $`\widehat{E}\left[F_t|F_0\right]=F_0`$ <sup>1</sup><sup>1</sup>1 To prove this, simply use the fact that the random variable $`X_t=_0^tW(u)𝑑u`$ with $`W(0)=0`$ is normally distributed with zero mean and variance $`\frac{t^3}{3}`$., it is in general not the case that $`\widehat{E}\left[F_t|F_u\right]=F_u,u>0`$, so the futures price is not a martingale: because $`\frac{S}{t}`$ persists for some time, observing the ‘momentum’ of the stock price would provide arbitrage opportunities between stock-returns and funding costs. In this model, options are priced at their discounted intrinsic value, which is consistent with the observation that holding $`\mathrm{\Gamma }`$ is of no value. In reality, people pay a price for the privilege of being long $`\mathrm{\Gamma }`$ and the $`\mathrm{\Theta }`$ is rather smaller than the cost of funding on the option premium and often negative, which thus indicates that the market believes that the stock price does have higher than linear variations. This means that the option market ‘knows’ that the stock price is either diffusive or exhibits jumps, or both. In turn, this also means that the arbitrage opportunity related to the persistence of the stock-momentum disappears. To state this formally: ###### Proposition 1 For the future $`F`$ of an asset $`S`$ (in a world of deterministic funding costs) which follows the Ito process $$dF=\left(\xi B(t)\frac{\xi ^2t^2}{2}\right)Fdt+\sigma (F,t)FdW$$ where $`B(t)`$ and $`W(t)`$ are SBMs, $`F,\xi ,\sigma 0`$, and the entire information on the asset price is it’s history, the following statements are equivalent 1. $`B(t)`$ is observable 2. $`\sigma (F,t)0`$ and $`\xi 0`$. 3. $`\xi 0`$, and the market satisfies, at any given time $`B(t)\stackrel{>}{<}\frac{\xi t^2}{2}dF\stackrel{>}{<}0`$. 4. The future price is predictable to order $`dt`$, and, except when $`B(t)=\frac{\xi t^2}{2}`$, allows arbitrage. 5. The price process $`F(t)`$ is stochastic and differentiable almost everywhere. It thus has no quadratic or higher variation. Also, if any of the conditions is true, there can be no risk-neutral measure for $`F`$. Proof. $`12`$ is trivial, as no two Gaussian variables can be simultaneously inferred from one futures price. $`23`$ follows from the fact that, if $`\sigma 0`$, almost all realisations of $`dW`$ would dominate over the $`dt`$-term in $`dF`$, so the statements in item 3 could not be made with any certainty, and vice versa. The first part of $`24`$ is trivial, since the leading order in $`dF`$ is $`dt`$ with coefficients which are (by equivalence with item 1) observable, if conditions $`2`$ are satisfied. This, however, would still be true if $`\xi `$ were zero. However, in this case $`dF0`$, and the market would not allow arbitrage. Yet if $`\xi 0`$, then we know the sign of $`dF`$, and arbitrage exists, in that a zero-price position in $`F`$ can be acquired that is instantaneously profitable. Thus also $`42`$. $`52`$ can be shown in the following way: if $`\sigma 0`$ anywhere, the price process would have quadratic variation, as there are terms of order $`\sqrt{dt}`$. If $`\sigma 0`$ and $`\xi 0`$, however, the process is deterministic. Thus a stochastic price process without quadratic variation must have $`\sigma 0`$ and $`\xi 0`$. $`25`$ is then just the trivial reversal of this argument. The non-existence of the risk-neutral measure follows directly from the existence of arbitrage opportunities, according to Harrison and Pliska $`\mathrm{}`$. Note also that the observation, originally made in the HJM framework for interest rates, that the no-arbitrage condition determines the drift of a stochastic, tradable variable is related to this proposition. This observations was extended to stochastic local volatilities by Derman and Kani . For both interest rates and stochastic volatilities it is also true that in the absence of quadratic or higher variations (i.e. zero volatility in a jump-less diffusion model) the no-arbitrage drift is zero. Proposition 1 only applies to a specific type of process, but makes no assumptions as to the nature of the tradable asset. Let us therefore assume the somewhat more general jump-less, stochastic futures price process $$dS(t)=\left(r(t)+I(t)\frac{}{t}\mathrm{ln}\widehat{E}_0\left[e^{_0^tI(u)𝑑u}\right]\right)S(t)dt+\sigma (W,s)S(t)dW(s),$$ (0.1) where $`I(t)`$ is an Ito process, and $`\widehat{E}_0`$ is the expectation operator evaluated at time zero. Henceforth, we call a process of the type Eq. (0.1) a generalised Ito process, which differs from the usual Ito process because the drift is an Ito process itself, but with a deterministic drift correction $`\frac{}{t}\mathrm{ln}\widehat{E}_0\left[e^{_0^tI(u)𝑑u}\right]`$. Funding rates are deterministic, such that the futures price $`F(t)`$ follows the same process, but without the interest rate drift component. This then has the solution $$F(t)=F_0\frac{\mathrm{exp}\left[_0^tI(s)𝑑s+_0^t\sigma (F(s),s)𝑑W\frac{1}{2}_0^t\sigma ^2(F(s),s)𝑑s\right]}{\widehat{E}\left[\mathrm{exp}_0^tI(s)𝑑s\right]},$$ which has similar properties to the process of Proposition 1, but is more general. This leads to a natural generalisation of Proposition 1: ###### Theorem 1 For a future – in a world of deterministic funding costs – on an asset, which follows the process Eq. (0.1), the following statements are equivalent 1. $`\sigma (F,t)0`$ and $`I(t)`$ is stochastic. 2. $`I(t)`$ is stochastic, and, at any given time $`I(t)\stackrel{>}{<}\frac{}{t}\mathrm{ln}\widehat{E}_0\left[\mathrm{exp}_0^tI(s)𝑑s\right]dF\stackrel{>}{<}0`$ 3. The future price is predictable to order $`dt`$, and, except when $`\frac{F}{t_{}}=0`$, allows arbitrage. 4. The price process $`F(t)`$ is stochastic without quadratic variation. Furthermore, if these conditions are true, there is no risk-neutral measure for $`F`$. Proof. As opposed to Proposition 1, here we have to leave out the observability of the realised value of $`I(t)`$, as this would require knowledge of $`\frac{}{t}\mathrm{ln}\widehat{E}_0\left[\mathrm{exp}^tI(s)𝑑s\right]`$, which can not be taken for granted. $`12`$ is analogous to Proposition 1, and items 3 and 4 follow immediately $`\mathrm{}`$. The important lesson to learn from Theorem 1 is that we do not need to know the details of the generalised Ito process that makes $`F`$ stochastic in order to conclude that in the absence of quadratic variation there are arbitrage opportunities which we can spot simply by observing $`\frac{F}{t}`$, and the only way to exclude arbitrage at all time is to make $`F(t)`$ deterministic and $`\widehat{E}_t\left[dF_u\right]=0u>t`$, which of course means that $`F(t)`$ is simply a constant in time. From this we introduce the important ###### Lemma 1 If, in a world of deterministic funding costs, a tradable asset follows a process with no quadratic or higher variation, it is either deterministic or allows arbitrage almost always, or both. Proof. The fact that the process has no higher than quadratic variations and that the futures price is a martingale means that it is a generalised Ito process of the type Eq. (0.1), except that the drift term $`I(t)`$ in Theorem 1 can be generalised to a jump-diffusion process where the set of times where jumps occur are of measure zero (i.e. the jumps are always countable, like in a Poisson process). Then the quadratic variation of the process $`F(t)`$ is zero, and the price-process is almost always predictable at least to order $`dt`$. This means it is differentiable almost everywhere, and continuous everywhere. The differentiability almost everywhere would allow successful arbitrage almost always (with infinitesimal losses in the countable cases where it is not differentiable), if the slope of the process is different from the funding costs on the asset (because the cost of setting up a futures position is zero). Therefore, the process allows no arbitrage only if $`F`$ is constant almost everywhere. Furthermore, if the generalised Ito process of no quadratic variation does allow arbitrage, it is very easy to spot the arbitrage by observing the momentum of the futures price $`\mathrm{}`$. It is very important to note that we never implied in any of the proofs of the above theorem and lemma that the future is written on a stock-price: It could have been a local volatility, as long as the funding costs are deterministic. Again, this is a generalisation of the observations of HJM, and Derman and Kani to any stochastic tradable asset under zero volatility . The HJM drifts for interest rates and the local-volatility drifts as calculated by Derman and Kani prove that even for non-zero volatility the drifts must be deterministic, if the no-arbitrage condition is to be satisfied. Lemma 1 is thus also more restricted than the results of , in that we only analyse the zero-volatility case, but it is also more general, in that it applies to any tradable asset. The bit to keep in mind for the next sections is that a stochastic tradable has to have at least quadratic variation in order to preclude arbitrage. ### 2.1 Does the quadratic variation matter in discrete hedging? Against taking quadratic variation too seriously, one can frequently hear variations of the following objection: ‘Quadratic variation has to do with the fractal dimension of the process. It basically says, if I add the moduli of all the stock-price changes in any finite interval, I get infinity. This is patently not the case in reality. Furthermore, I typically re-hedge only after a stock-move of, say, 4%. So I re-hedge maybe once a week on average. Adding all the moduli of stock-returns relevant to my hedging-strategy, they obviously don’t diverge. So for practical purposes, the stock-movement relevant to me is proportional to time. In other words: to me the stock-price path has dimensionality 1’. It is easy to see what is wrong with this argument. Our hedger experiences a typical $`\mathrm{\Gamma }`$-related P&L component of $`52(0.04)^2\mathrm{\Gamma }`$ on an annualised basis. The crux is that, if his pain-barrier was a stock-move of, say, 2%, he would not re-hedge once a week, or even twice a week on average, but four times a week, if the quadratic variation of the stock-price process is the highest non-vanishing variation. So on an annualised basis, his $`\mathrm{\Gamma }`$-related P&L component is $`208(0.02)^2\mathrm{\Gamma }`$, which is exactly the same as before. The quadratic variation therefore does matter equally for someone who hedges infrequently as for someone who hedges continuously. One can not escape the $`\mathrm{\Gamma }`$-related P&L by simply re-hedging more frequently. In other words: the fractal dimension of the stock-price process matters regardless of the granularity of price observations. If our hedger lived in a world where the dimensionality of the stock-price process is linear, dropping his pain-barrier to 2% would incur a $`\mathrm{\Gamma }`$-related P&L component of $`104(0.04)^2\mathrm{\Gamma }`$, and he could observe, regardless of the granularity of his hedge, that more frequent re-hedging lets him reduce the $`\mathrm{\Gamma }`$-P&L. Let us therefore not be fooled: The fractal dimension of the stochastic process is observable. In particular: ###### Theorem 2 The fractal dimension on the scale relevant to discrete-time hedging is also observable. ### 2.2 Hedging ‘Exotics’ in the Fixed-Income World Before we tackle the stochastic local volatility world, let us consider an instructive example of the typical problem we discussed in section 2 from the fixed-income markets. Kani et. al. observed that trading and hedging against forward interest rates is similar to trading and hedging against local volatilities . This allows us to use examples which are easy to follow because the set of independent forward rates is one-dimensional, while the set of local volatilities is two-dimensional. Here we bring an example of a very simple fixed income world with only two bond-prices, and a simple ‘exotic’ (i.e. a simple function of these two prices) which can not be replicated with a dynamic trading strategy without taking new independent variables into account, namely variance and covariance rates of the interest rates. This is a further qualificatnion of the claim by Kani et. al., namely that everything in the fixed-income world can be replicated by dynamically attainable ‘forward rate gadgets’ which have zero cost and provide a specified degree of exposure to a particular forward rate . The central conclusion we work towards here is the inapplicability of an effective DLVF theory to an SLVF world. Consider thus the frequent claim that vanilla options make local-volatility gadgets attainable, which complete the market even in the presence of exotic options. This claim and the above-mentioned claim of Kani et. al. are are easily mis-interpreted for the same reason, namely that quadratic variations are not negligible in a stochastic, arbitrage-free market, and that they can not be fully hedged against <sup>2</sup><sup>2</sup>2We would like to stress that ref. indicates that Kani et. al. are not vicitims of this mis-interpretation. What matters here is only that ref. could easily mislead the reader into believing that forward-rate gadgets are perfect hedging instruments, as the paper does not refer to the necessity of making convexity corrections. The point we make is that misreading the paper in this way is analogous to ignoring the quadratic variation of local volatility in DLVF models.. Note, however, that this observation can already be made in the derivation of the Black-Scholes model: Perfect replication of a vanilla option in the Black-Scholes world is only possible when we also know the volatility of the stock. For multi-factor models, we need to know the variance-covariance rate matrix in order to achieve perfect replication. The DLVF models make no reference to this matrix, and there are usually not enough degrees of freedom to calibrate to such a matrix if it were known. Consider a world with only two bonds $`B_1`$ and $`B_2`$, maturing at times $`t_1`$ and $`t_2`$ from now. The corresponding interest rates are defined through $`B_1=\mathrm{exp}(r_1t_1)`$ and $`B_2=\mathrm{exp}(r_2t_2)`$. The forward rate is the interest rate for the term between $`t_1`$ and $`t_2`$ which can be locked in at present, i.e. $`f=\frac{r_2t_2r_1t_1}{t_2t_1}`$, therefore $`\frac{B_1}{B_2}=\mathrm{exp}\left(f(t_2t_1)\right)`$. The ‘interest rate gadget’ is the portfolio $`\mathrm{\Lambda }=B_1\left[\frac{B_1}{B_2}\right]B_2`$, where the square brackets indicate a position size (‘hedge ratio’), which means a value that stays fixed during instantaneous changes of the market observables $`r_1`$ and $`r_2`$. Now imagine we have some other instrument, whose exposure to the forward rate $`f`$ we want to hedge. We observe that $$\delta \mathrm{\Lambda }=B_1\left[(t_2t_1)\delta ft_1(t_2t_1)\delta f\delta r_1\frac{1}{2}(t_2t_1)^2(\delta f)^2\right].$$ Kani et. al. mention that the bond $`B_1`$ can be synthesised by holding one gadget and $`\frac{B_1}{B_2}`$ units of $`B_2`$, which gives exact replication to all orders . Unfortunately, this is not a valid example of hedging, because all we have in the ‘replicating’ portfolio is exactly one bond $`B_1`$. To really keep the correspondence with DLVF models, we need to introduce a new instrument which has a pay-off as a function of $`f`$, but which is not already in this market. Consider for example a security which pays exactly the forward rate as, say, a US dollar amount. To hedge this, we need a strategy whose P&L is and stays linear in the forward rate. However, since the hedge ratio itself is a function of the forward rate, we get the second derivatives in $`\delta \mathrm{\Lambda }`$. If we hedge the forward-rate exposure of some portfolio with the gadget $`\mathrm{\Lambda }`$, we see that we introduce a negative P&L component proportional to the variance rate of $`f`$, and a correlation term between the forward rate and $`r_1`$. We do not have enough degrees of freedom to eliminate these new terms, i.e. to synthesise a product which is strictly linear in the forward rate. The corresponding effect on, say, for example, interest rate futures prices is frequently referred to in the fixed income markets as a ‘convexity correction’. The size of the convexity correction is increasing with the variance rate $`(df)^2/dt`$ and the covariance rate $`(df)(dr_1)/dt`$. Note that if we know both these rates, perfect replication is still possible, in the same sense as it is possible in the Black-Scholes world for vanilla options when we know the volatility of the underlying. ## 3 Stochastic Local Volatilities Define $`\stackrel{}{x}=(F_S,\stackrel{}{\alpha })`$, where $`\stackrel{}{\alpha }`$ is an $`n`$-component vector parameterising the local volatility surface, calibrated to market prices, and $`F_S`$ is the forward price of the stock. Let us now consider a scalar function $`f(\stackrel{}{x})`$ of these $`n+1`$ stochastic variables, and define the vector-operator $$\stackrel{}{}(\frac{}{F_S},\frac{}{\alpha _1},\frac{}{\alpha _2},\mathrm{},\frac{}{\alpha _n}).$$ Then $$df=f_tdt+(\stackrel{}{}f)(d\stackrel{}{x})+\frac{1}{2}(d\stackrel{}{x})^T(\stackrel{}{}\stackrel{}{}f)(d\stackrel{}{x})+\mathrm{}$$ If the stochastic process for $`\stackrel{}{x}`$ has no more than quadratic variation, the higher order terms are smaller than order $`dt`$. Let us assume that local volatility futures (or local volatility ‘gadgets’ in the language of ref. ) are attainable at zero cost. Then, after hedging the portfolio to first order in the stock price and local volatilities, the continuous trading equation is $$f_tdt+\frac{1}{2}(d\stackrel{}{x})^T\left(\stackrel{}{}\stackrel{}{}f\right)(d\stackrel{}{x})=rfdt,$$ (0.2) which makes for a total of one $`\mathrm{\Theta }`$ component, the funding on the stock-hedge, and $`\frac{(n+1)n}{2}`$ terms arising from all the different second derivatives with respect to the different components of $`\stackrel{}{x}`$ <sup>3</sup><sup>3</sup>3 The assumption that local volatility futures are attainable is probably no more permissible than the assumption in the fixed income world that forward rate gadgets are attainable, that is they are attainable only to first order. In this sense, Eq. (0.2) will not be strictly correct, but will contain other second order terms, arising from the second-order terms of the local volatility gadgets. Qualitatively, the number of P&L terms is predicted correctly, however, as we only need replace $`\left(f\right)_{ij}`$ with $`\left(f\right)_{ij}+C_{ij}`$, where $`C_{ij}`$ represents the combined second order exposure created by our holding the local volatility gadgets.. We now have to make assumptions on what the asset price process is in the real world. Broadly, we can classify the imaginable price processes into four groups: * DLVF world: The local volatilities are completely deterministic. The Black-Scholes world is a trivial example. The asset price process of an implied tree is the most general case. * WSLVF world: The local volatilities are stochastic, continuous everywhere, and differentiable almost everywhere. Let us call this model the ‘weakly stochastic local volatility function’ model. * SLVF world: The variables parameterising the volatility function follow stochastic processes with non-vanishing quadratic variation, but vanishing higher variations (This means these processes are diffusive). Call this the stochastic local volatility function model. * Jump-diffusion world: Either the stock-price process or the volatility parameters (as implied by the vanilla options market) or both are not simply diffusive. This would manifest itself in higher than quadratic variations. Note that the decision on which world a particular market corresponds to can be based on empirical data according to Theorem 2. Let us consider the suitability of the DLVF model to all of these four worlds. In the DLVF world the local volatilities are perfect forecasts of future realised instantaneous volatilities. In the absence of arbitrage, the evolution of the local volatility surface is static except for a shift of the time-frame of the observer. In this case, risk-neutral valuation and the DLVF model are obviously applicable. However, the viewpoint that we live in such a DLVF world contradicts empirical evidence . The WSLVF world has the previously described arbitrage opportunities: observe the local volatility ‘momentum’, and simply go long the gadgets corresponding to growing local vols, and short the others. This provides a risk-less profit almost always <sup>4</sup><sup>4</sup>4 A necessary condition for the existence of such arbitrage is that the ‘almost never’ occurring cases of making a loss produce a bounded loss, which is the case here.. Within this world the implied tree does not work in the way it is customarily used and clearly all vanilla options and all options calculated on the implied tree are priced assuming a sub-optimal hedging strategy. In fact there is no optimal hedging strategy: any given strategy can be beaten by another one using more of the arbitrage opportunities. Clearly, while the world could be WSLVF, we gain nothing by making this assumption, unless we have a model to identify and exploit the arbitrage opportunities. We can make the intriguing observation, however, that everyone gets away with using risk-neutral valuation as long as no market participant spots the arbitrage. If option buyers and sellers all worked with a DLVF model and the associated hedging strategies, the existence of arbitrage is irrelevant for practical purposes. So it is in fact entirely possible that the real world is of the WSLVF type, although, we would clearly prefer to have a stronger cause for the use of risk-neutral valuation in the DLVF model than the reason that it works simply because everybody is using it. Note that the assumption that the real world is WSLVF is easily testable: use a DLVF model, and take local volatility positions according to the ‘momentum’ of the local volatility parameters. Unless one makes profits almost always with this strategy, the real world is not WSLVF. Next we have to discuss the SLVF world. A necessary condition for local volatility gadgets to exist is that tradability of all European vanilla options, which is equivalent with the tradability of all Arrow-Debreu options. Since the existence of sufficiently many of these options is a necessary condition to calibrate the LVF surface, we will assume in the remaining part of this paper that sufficiently many Arrow-Debreu options or European vanilla options are available in order to produce gadgets for all the LVF parameters we allow in the model, and that these options can be used in any hedging strategy for exotic options. Then we have two possibilities to distinguish: * SLVF(I) world: Local volatility gadgets exist, but second-order exposure can not be hedged. The gadgets themselves may or may not produce additional second-order exposure. * SLVF(II) world: In this world, we can hedge against variations of stock price and local volatilities up to second order. The central issue of this paper is that the SLVF(II) world is theoretically impossible unless there is a very low-parametric representation of the LVF, and that, in any case, volatility of volatility has P&L effects which accrue over time. We do not consider the jump-diffusion world for two reasons. Firstly, we are not aware of any specific claim that DLVF models could be appropriate to such a world, although the idea that effective theories may also have jump-risk ‘effectively’ integrated out can not be discarded a priori. Secondly, however, we already show that there irreconcilable problems with the application of DLVF models in an SLVF world. Since jump-diffusion models are a super-set of SLVF models, we would have to hedge against even more sources of uncertainty in jump-diffusion worlds. It is thus unnecessary to prove that the problems of the SLVF worlds persist in the jump-diffusion setting. ## 4 P&L in the SLVF worlds In an SLVF(I) world, we have all the P&L terms of Eq. (0.2) to consider. This will lead to risk premia on the option proportional to the variance of the to-be-realised ‘$`\mathrm{\Gamma }`$’ P&L, in other words proportional to $$d\stackrel{}{x}^T_0^T\left(f\right)𝑑\stackrel{}{x},$$ (0.3) or simply to all the un-hedged second order terms, because they are of order $`dt`$ and thus accrue P&L proportional with time. The question is now whether we can quantify the volatility risk premium appropriate for an exotic option in an SLVF(I) model. Obviously, the terms in Eq. (0.3) are in principle computable, if the market-implied covariances between all local volatilities are known, as the rest is simply path-integration <sup>5</sup><sup>5</sup>5 We make no statement about the practical difficulties of attempting this computation.. There is of course no hope to infer these parameters unless the a priori parameterisation of the LVF we are willing to consider is extremely low-dimensional: For an $`n`$-dimensional parameterisation, we have to infer the $`n`$ parameters, and all components of the variance-covariance matrix of these parameters plus the stock-price. This is a symmetric $`(n+1)\times (n+1)`$ matrix. In total, we have thus $`\frac{(n+1)(n+2)}{2}+n`$ parameters to estimate (in a non-mean-reverting SLVF model). Assuming we have $`m`$ independent vanilla option prices, we are allowed to use no more than $`\frac{\sqrt{41+8m}7}{2}`$ parameters for the LVF. We thus need four vanilla option prices for the usual non-mean-reverting stochastic volatility model, and at least eight option prices to accommodate even the simplest skew information, by specifying the SLVF by two volatility parameters. If the SLVF is required to be mean reverting, we require twelve options to calibrate such a model. Now say one of those parameters was designed to capture skew, and the other to capture term structure. If we therefore took a second skew parameter, making the total parameterisation of the LVF three-dimensional, we would require 19 options in the mean-reverting case and 13 otherwise. This analysis underlines the urgent need for ‘smoothing’ or ‘regularising’ of the LVF, which has been attempted by many authors recently . However, smoothing alone does not suffice to estimate volatility risk premia. At this point in the discussion, one might want to ask the question of whether the volatility risk premium really matters. After all, we never worry about it in the Black-Scholes world, even though the Black-Scholes assumption of deterministic volatility is clearly wrong in practice. ### 4.1 Volatility Risk Premia and Exotic Options It is in the context of Black-Scholes type models, that we find that there is skew: The BS volatilities implied in the market differ for different strikes and maturities. This is ascribed to the BS assumptions being wrong in one way or another. The LVF approach is to say that volatility simply is not flat, the stochastic volatility and jump-diffusion approach is to say that there is risk-aversion and risk-premia are associated with un-hedgeable risks, or in other words with the incompleteness of the market. When using a BS-implied volatility surface, we do not need to worry what causes the skew. If there are risk-premia, they will already be in the option price. However, there is a qualifying statement to be made: the P&L of a $`\mathrm{\Delta }`$-hedged long option position until expiry is $$_0^T\left(\mathrm{\Theta }(t)+\frac{1}{2}\mathrm{\Gamma }_F(t)F^2(t)\widehat{\sigma }^2(t)r(t)f(t)\right)𝑑t$$ (0.4) where $`\widehat{\sigma }(t)`$ is the actual realised volatility at time $`t`$, and $`\mathrm{\Gamma }_F`$ is the futures-$`\mathrm{\Gamma }`$. All the integrands are stochastic, and we thus do not expect the option to be priced at the expectation of all these parameters. Instead, we expect that the option is priced at the expectation – under some risk-neutral or, if necessary, risk-adjusted measure – of the entire integral Eq. (0.4). This price can be expressed by means of a Black-Scholes implied volatility. Thus, the option can be priced at some implied $`\sigma `$ which can, in Eq. (0.4), replace the stochastic $`\widehat{\sigma }(t)`$. The Black-Scholes implied volatility allows us to use the Black-Scholes option pricing theory, which is an effective theory, where specific expectations (or integrals) of some of the underlying stochastic variables are extracted from the current prices of the traded assets. The effective dynamics which results is based on some of the sources of uncertainty being ‘effectively’ integrated out of the full stochastic theory. It can be shown that DLVFs are risk-adjusted expectations of future instantaneous volatilities . In this way, DLVF models are effective theories of volatility in the same way as static forward interest rates define an effective theory for interest rates. This is an important point which we will return to later on. Beyond the volatility risk in Eq. (0.4) of the P&L of a hedged options position, there is an equally important risk of spending an unusually long amount of time in regions of high $`\mathrm{\Gamma }`$. However, we should expect that no-one is compensated for this option-specific risk, simply because this can be diversified away by selling a portfolio of vanilla options providing a reasonably constant $`\mathrm{\Gamma }`$ profile across spot prices and times. This portfolio would still have vega exposure, but no option-specific risk <sup>6</sup><sup>6</sup>6 In fact there is some option-specific risk, because when the implied vols change for some options, the $`\mathrm{\Gamma }`$-profile of the portfolio changes. This is also a vega risk, because it is caused by the change of implied volatilities.. We thus formulate the following observations. ###### 1 The implied Black-Scholes volatilities compensate the option seller for volatility risk, but not for option-specific risk. If we use implied Black-Scholes volatilities to price exotic options, we ignore that the exotic may have specific risk that can not be diversified away by trading listed options. In particular that risk, because it is a specific risk attached to an un-quoted asset, can not be implied from the market. Furthermore, the same argument holds for any model calibrated only on vanilla options. This leads to the key criterion for successful exotic model valuation: ###### 2 An option for which there is no known market price must either be hedgeable with quoted securities, or otherwise will incur instrument-specific risk premia. Obviously, this condition is satisfied by any sensible model whenever reality stays within the constraints of the model-assumptions. The problems with the DLVF models, as we argue here, is that reality is too far removed from the fundamental DLVF assumption of deterministic (local) volatility parameters. This is underlined by empirical research on market data . A trivial aside of the above is ###### 3 An exotic option model has to either have a plausible way of incorporating option-specific risk-premia or permit a proof that a replicating trading strategy exists. We will show below that the DLVF model applied to a world of non-deterministic volatility parameters does neither if the only allowed hedging and calibration instruments are European vanilla options. It does not matter whether we use a low or high-parametric SLVF model: In low-parametric SLVF models we may create option-specific risk that can not be hedged by other vanilla options, simply because the parameter-family of LVFs do not capture the full volatility exposure of all conceivable exotics. In a high-parametric SLVF, we will normally not have enough vanilla options to calibrate to, and worse: even if we have a continuum of vanilla options for all strikes and expiries, we can not calibrate a continuum of SLVF parameters to that. It is therefore relevant to initially construct the space of all options we are willing to consider as hedging instruments in our model, and then decide which SLVF world this puts us in. Obviously, in an SLVF(II) world, there always is a hedging strategy for everything, but it would have to be shown that we are in such a world. The next step is thus to provide a mechanism which proves the existence of a hedge against stock and volatility risk in an LVF framework. We choose a trinomial world. More generally, we simply try to answer the question of which of the SLVF worlds must be the real one, if the real asset price process is SLVF, and if the set of all exotics we wish to consider are all exotics that can be defined in this discrete world. It therefore leads naturally to an answer to the question whether all well-defined exotics can be hedged in the continuum limit, and whether this continuum limit is an SLVF(I) or SLVF(II) world if we only consider vanilla options in the hedge. What is important to remember, however, is that we will perform the model calibration to the vanilla European options market in the DLVF sense, i.e. we do not calibrate the local volatilities’ variance and covariance rates. This is because we want to test the viability of the DLVF model in a world where the asset price process is SLVF, and where only vanilla options, the underlying, and funding instruments are considered as hedging instruments. ## 5 Hedging an Exotic Option in an SLVF world using a DLVF model ### 5.1 Geometry of the tree A discrete local volatility tree with pre-determined state space needs to be at least trinomial in order for the discrete process to recombine <sup>7</sup><sup>7</sup>7 Implied binomial trees are possible, but the state space (i.e. the up and down steps and the time steps) depends on synthetic options which are very short-dated and forward starting a long time in the future. We prefer not to import the complications of the market-calibration into the construction of the state-space, so that volatility changes leave the state-space unchanged. Also, the trinomial approach is much simpler, as the Eqs. (0.5) leave exactly one free local parameter: the local volatility.. Since we assume that interest rates can be hedged separately, we work in a zero-interest-rate tree (the appropriate change of numeraire is thus implicitly assumed), which we can specify with the following state-space and transition probabilities $`S_n^i=S_0e^{iu},`$ $`nin`$ $`p\left(S_n^iS_{n+1}^{i1}\right)`$ $`={\displaystyle \frac{\alpha _n^i}{1+e^u}}`$ $`p\left(S_n^iS_{n+1}^i\right)`$ $`=1\alpha _n^i`$ $`p\left(S_n^iS_{n+1}^{i+1}\right)`$ $`={\displaystyle \frac{\alpha _n^i}{1+e^u}}`$ (0.5) where $`S_0`$ is the initial value of the stock price. At any step, the stock price change by a factor $`e^{\pm u}`$ or stay the same. $`S_n^i`$ is the stock price at the $`n^{\mathrm{th}}`$ non-trivial time-slice after $`i`$ net up-moves. The choice of notation for the transition probabilities is an arbitrary parameterisation such that the sum of the probabilities to leave a given node is one, and the stock price is a martingale, and the $`\alpha `$’s are strictly monotonically increasing functions of the local volatilities. The choice of the ‘global’ tree-parameter $`u`$ puts an upper bound on the possible local volatilities in this world, because, in order to preserve positivity of the transition probabilities, we require $`0\alpha 1`$. This does not put any fundamental restrictions on the conclusions we will draw from analysing this discrete model <sup>8</sup><sup>8</sup>8 From $`\widehat{E}\left[\frac{S_{t+1}}{S_t}\right]=1`$ and $`\widehat{E}\left[\left(\frac{S_{t+1}}{S_t}1\right)^2\right]=e^{\sigma ^2\mathrm{\Delta }t}1`$ we get $`\alpha _\tau ^i=e^u\frac{e^{\left(\sigma _\tau ^i\right)^2\mathrm{\Delta }t}1}{\left(e^u1\right)^2}`$, and the constraint $`\alpha _\tau ^i1`$ means that all local volatilities have to satisfy $`\left(\sigma _\tau ^i\right)^2\frac{1}{\mathrm{\Delta }t}\mathrm{ln}(2\mathrm{cosh}u1)`$.. We make no assumptions about the process underlying the evolution of the local volatility parameters $`\alpha _n^i`$. For example, restrictions on the drift of local volatilities have to be posed by the condition that, if local-volatility gadgets are attainable, the local volatilities and the stock price have to be jointly martingale in order to avoid arbitrage . However, we do not need to place any restrictions on the evolution of the volatilities, except that they are bounded in the interval $`[0,u]`$ <sup>9</sup><sup>9</sup>9 A necessary condition for this is that the volatilities of local volatilities are square-integrable over time and that the modulus of the drift are integrable over time, although these two conditions are not sufficient.. ### 5.2 Paths and vanilla options on the tree The set of all paths $`𝒫`$ has $`3^n`$ elements for a tree with $`n`$ time-slices. Define a path as the $`n`$-element sequence of the space-slices the path visits $$\pi =(\pi _1,\pi _2,\mathrm{},\pi _n),\left(\pi _0=0,\left(\pi _i\pi _{i1}\right)\{1,0,1\}\right).$$ We will also use the alternative definition, the $`n`$ element sequence of steps (1=up, -1=down) $$\stackrel{~}{\pi }=\{\stackrel{~}{\pi }_1,\stackrel{~}{\pi }_2,\mathrm{},\stackrel{~}{\pi }_n\},\stackrel{~}{\pi }_i\{1,0,1\}.$$ The realised stock-path $`\mathrm{\Pi }(m)`$ up to time $`m`$ gives a filtration on the set of paths $`𝒫`$. We define $`𝒫\left[\mathrm{\Pi }(m)\right]𝒫`$ as the set of the paths whose first $`m`$ steps match the realised path $`\mathrm{\Pi }(m)`$, and the final stock-value of the realised path as $`S_0e^{\mathrm{\Pi }_m(m)u}`$. We also define the set of directed paths $`𝒟[m,i,\tau ,j]=\left\{\pi 𝒫,\pi _m=i,\pi _\tau =j\right\}`$, which means $`𝒟[m,i,\tau ,j]`$ is the set of all paths which at time $`m`$ give a stock price of $`S_0e^{iu}`$ and at time $`\tau `$ a stock price of $`S_0e^{ju}`$. Let us define the future cone of a node as the part of the tree which can be reached from that particular node. The question is whether it is possible to hedge any path-dependent derivative when, at any time-step $`m\{0,1,\mathrm{},n1\}`$, the local volatilities $`\alpha _k^i`$ in the future cone of $`(m,\mathrm{\Pi }(m))`$ are not known with certainty, and we only use European vanilla options as hedge instruments? The restriction to use only European options is equivalent to using only Arrow-Debreu options. On the lattice-world, this is trivially true, whereas in the continuous world, some limiting procedure is necessary. The path-probabilities, as measured from time $`m`$, are $`p`$ $`(\pi ,m,\mathrm{\Pi }(m))=𝐈_{\pi 𝒫[\mathrm{\Pi }(m)]}\times `$ $`{\displaystyle \underset{j=m}{\overset{n1}{}}}\left(𝐈_{\stackrel{~}{\pi }_{j+1}=1}{\displaystyle \frac{\alpha _j^{\pi _j}(m)}{1+e^u}}+𝐈_{\stackrel{~}{\pi }_{j+1}=0}\left(1\alpha _j^{\pi _j}(m)\right)+𝐈_{\stackrel{~}{\pi }_{j+1}=1}{\displaystyle \frac{\alpha _j^{\pi _j}(m)}{1+e^u}}\right)`$ where $`𝐈_x`$ is the indicator function which is equal to 1 if $`x`$ is true and zero otherwise. The Arrow Debreu price $`A_\tau ^i(m)`$ is the price at time $`m`$ of the derivative paying one unit if the stock-price at time $`\tau `$ is $`S_0e^{iu}`$, i.e. if the future realised path satisfies $`\mathrm{\Pi }(\tau )𝒫\left[\mathrm{\Pi }(m)\right]𝒟[0,0,\tau ,j]`$. Since $`p(\pi ,m,\mathrm{\Pi }(m))`$ is zero whenever $`\pi 𝒫\left[\mathrm{\Pi }(m)\right]`$, $$A_\tau ^i(m,\mathrm{\Pi }(m))=\underset{\pi 𝒟[0,0,\tau ,j]}{}p(\pi ,m,\mathrm{\Pi }(m)),\tau >m.$$ Note that the continuum limit of a discrete Arrow-Debreu price is an ill-defined concept. Instead, a given continuous-world definition of an Arrow-Debreu price needs to be matched to a discretisation-dependent linear combination of discrete-world Arrow-Debreu prices. This will then converge properly to the discrete-world Arrow-Debreu price. This is normal procedure for path-integration problems, and is implicitly done for any tree-model. ### 5.3 Path-dependent options The most general path dependent option $`f`$ has a final pay-out of $`X=X\left(\mathrm{\Pi }(n)\right)`$. In a similar fashion as for the Arrow-Debreu options, we need to ensure that this definition provides a correct continuum limit for the path integral. Thus, for the option to approximate a continuum-world exotic, $`X=X\left(\mathrm{\Pi }(n)\right)`$ is a discretisation-dependent function. The question is whether such a path-dependent option can always be hedged with a strategy involving only the stock and Arrow-Debreu options, or whether such a hedge becomes valid at least in the continuum limit. Under risk-neutral valuation, the value of the path-dependent option at time $`m`$ is $$f\left(\mathrm{\Pi }(m)\right)=\underset{\pi 𝒫}{}p(\pi ,m,\mathrm{\Pi }(m))X(\pi ).$$ To determine the hedge portfolio, we have to know how many independent securities there are available in this market. There are two constraints on the Arrow Debreu prices $$\underset{i=\tau }{\overset{\tau }{}}A_\tau ^i(m,\mathrm{\Pi }(m))=1$$ $$\underset{i=\tau }{\overset{\tau }{}}A_\tau ^i(m,\mathrm{\Pi }(m))e^{iu}=e^{\mathrm{\Pi }_m(m)u},$$ (0.6) valid for all $`\tau ,m;\tau >m`$. The first condition derives from the fact that the stock-price at time $`\tau `$ will with certainty be on one of the nodes of that time-slice, and the second simply says that a portfolio of Arrow Debreu prices which pays exactly the stock-price at any node at time $`\tau `$ is equivalent to a portfolio holding only the stock. Henceforth, for notational simplicity, we shall simply write $`\mathrm{\Pi }(m)`$ instead of $`(m,\mathrm{\Pi }(m))`$, as the time can be inferred from the cardinality of the representation of $`\mathrm{\Pi }(m)`$. At time $`m`$, the most general portfolio $`\mathrm{\Lambda }`$ short $`f`$ can thus be represented as $$\mathrm{\Lambda }(\mathrm{\Pi }(m))=f(\mathrm{\Pi }(m))+\omega (\mathrm{\Pi }(m))S(\mathrm{\Pi }(m))+\text{ }$$ $$\underset{t=m+1}{\overset{n}{}}\left(\underset{i=\mathrm{\Pi }_m(m)(\tau m)+2}{\overset{\mathrm{\Pi }_m(m)+(\tau m)}{}}w_\tau ^i(\mathrm{\Pi }(m))A_\tau ^i(\mathrm{\Pi }(m))\right).$$ (0.7) Note that the somewhat strange limits in the inner sum are there in order to avoid specifying weights for worthless and redundant Arrow Debreu options. Specifying the entire time-indexed functionals $`\omega (\mathrm{\Pi }(m))`$ and $`w_\tau ^i(\mathrm{\Pi }(m))`$ for all $`m`$ and $`\mathrm{\Pi }(m)`$ then defines a trading strategy. If there exists a hedged trading strategy, it must be completely determined by the specification of these functional. Note that we assumed we are in a zero-interest-rate environment, so we have disregarded the bond/deposit/loan in our portfolio, and the constraint that the portfolio should be self-financing does not apply in the usual sense. Instead, all we require from a hedged strategy is $$\mathrm{\Lambda }(\mathrm{\Pi }(m))=\mathrm{\Lambda }^{}(\mathrm{\Pi }(m+1))\mathrm{\Pi }(m+1)𝒫\left[\mathrm{\Pi }(m)\right],$$ where $`\mathrm{\Lambda }^{}`$ is the portfolio one time-step ahead, but before re-hedging. $`\mathrm{\Lambda }^{}(\mathrm{\Pi }(m+1))`$ thus has the same portfolio weights as $`\mathrm{\Lambda }(\mathrm{\Pi }(m))`$. As time moves on one step, the following relevant market parameters will have changed: * the stock price * the local volatilities in the future cone of $`\mathrm{\Pi }_m(m)`$. Given that we assume that we know the volatility $`\alpha _m^{\mathrm{\Pi }_m(m)}`$ with certainty (and given that, after $`\mathrm{\Delta }`$ and $`\mathrm{\Gamma }`$-hedging in the trinomial world we are in fact not exposed to it at all), we can work out how many constraints and free variables there are to work out the hedge at every step. ### 5.4 Constructing the hedge; tractor options Any path-dependent option (and therefore any option) can be written as a linear combination of what we shall call tractor options. A tractor options $`𝒯_\pi `$ pays out one unit if and only if the stock-price up to time $`n`$ has followed the path $`\pi `$. In the continuous world, we could call the payoff of any contingent claim (exotic or otherwise) a functional on the space of continuous functions $`S(t)`$. Each option valuation can be expressed as an infinite-dimensional integral over the set of continuous functions on the time to maturity $`C\left([0,T]\right)`$, with the corresponding pay-off functional as the integral kernel, which leads to the path-integral representation of risk-neutral option valuation. It is possible that the analysis which follows can be consistently formulated in this continuous framework, but the path-integral can also be represented as a limiting case of path-integrals over functions on discrete image spaces as long as the mapping from the continuous-world derivative to the corresponding discrete-world linear combination of tractor options is done in a way that demonstrably converges to the continuous-world pay-off. We are going this route of replacing the path-integral by a sum of paths in a discrete world, and then taking the continuum limit. The question of whether all exotic options can be hedged can now be re-phrased as the question of whether all tractor options can be hedged. Note that the value of a tractor option at time $`m`$ is simply $$𝒯_\pi \left(\mathrm{\Pi }(m)\right)=p(\pi ,\mathrm{\Pi }(m)),$$ so we see that we have in fact been using tractor options already in the discussion so far, and that – in the language of path-integration – the tractor option prices form the probability measure on the function space for the functions $`S(t)`$. So we see that $`𝒫`$ is the set of all states of the world, and the set of all tractors generates the largest possible $`\sigma `$-algebra on that space. $`𝒯:𝒫[0,1]`$ is our probability measure on that space, and $`(𝒫,\sigma (\left\{𝒯_{\pi _1}\right\},\left\{𝒯_{\pi _2}\right\},\mathrm{},\left\{𝒯_{\pi _{3^n}}\right\}),𝒯)`$ is the relevant probability space for valuing options in a risk-neutral valuation framework. It can also be seen that our above definition of an Arrow-Debreu price is itself a linear combination of tractor prices $$A_\tau ^i(\mathrm{\Pi }(m))=\underset{\pi 𝒟[0,0,\tau ,j]}{}𝒯_\pi (\mathrm{\Pi }(m)),\tau >m,$$ which means that the Arrow Debreu prices are the probabilities associated with a particular partition of the state space. Note that this equation holds independently of the local volatilities. We may ask whether the reverse is true, i.e. whether we can write the tractor as a linear combination of Arrow-Debreu options. The answer is strictly no, as can be easily seen by the fact that there are many more tractors than Arrow-Debreu prices, and there are probably no more constraints on the tractor prices than we had for the Arrow-Debreu prices in Eq. (0.6). So there is no general static hedging strategy possible for all tractors. In set-theoretical notation, the $`\sigma `$-algebra generated by the sets $`𝒟[0,0,\tau ,j]𝒫\left[\mathrm{\Pi }(m)\right]`$ (relevant to observing which Arrow-Debreu option pays off) is much smaller than the $`\sigma `$-algebra generated by the sets $`\{\pi _i\},\pi _i𝒫\left[\mathrm{\Pi }(m)\right]`$ (relevant to observing which tractor option pays off). Let us again consider the portfolio Eq. (0.7). How many free parameters are there to adjust the hedge with? It is easy to convince oneself that there will be $`(nm)^2`$ independent Arrow-Debreu weights <sup>10</sup><sup>10</sup>10 Since $`_{\tau =m+1}^n\left(_{i=(\tau m)+2}^{\tau m}1\right)=(nm)^2`$. and one stock-weight. The number of constraints depends on our hedging ambitions. We could attempt to hedge against the volatilities to any order, which we will demonstrate to be impossible for any tree of more than two steps. However, this is only necessary in a jump-diffusion world. If we want to be hedged against volatility to second order, as we should be in an SLVF world, the constraints are $`\mathrm{\Lambda }^{}\left(\mathrm{\Pi }(m)\{1\}\right)=\mathrm{\Lambda }^{}\left(\mathrm{\Pi }(m)\{0\}\right)=\mathrm{\Lambda }^{}\left(\mathrm{\Pi }(m)\{1\}\right)`$ $`{\displaystyle \frac{\mathrm{\Lambda }^{}\left(\mathrm{\Pi }(m)\{1\}\right)}{\alpha _i^j}}={\displaystyle \frac{\mathrm{\Lambda }^{}\left(\mathrm{\Pi }(m)\{1\}\right)}{\alpha _i^j}}`$ $`{\displaystyle \frac{\mathrm{\Lambda }^{}\left(\mathrm{\Pi }(m)\{0\}\right)}{\alpha _i^j}}=0`$ $`{\displaystyle \frac{^2\mathrm{\Lambda }^{}\left(\mathrm{\Pi }(m)\{0\}\right)}{\alpha _i^j\alpha _k^l}}=0`$ $`i=m+1,\mathrm{},n1,j=\mathrm{\Pi }_m(m)(im),\mathrm{},\mathrm{\Pi }_m(m)+(im)`$ $`k=m+1,\mathrm{},n1,j=\mathrm{\Pi }_m(m)(km),\mathrm{},\mathrm{\Pi }_m(m)+(km)`$ (0.8) The first equation enforces $`\mathrm{\Delta }`$ and $`\mathrm{\Gamma }`$ hedging, the second equation enforces cross-terms between volatility and stock, the third equation enforces ‘local-vol-parameter’ vega hedging, and the fourth enforces all the second derivatives with respect to the volatilities’ parameterisation. Note that in the second line in Eq. (0.8) we do not require the first derivatives to be zero, as we do in the third line, as this would also enforce a hedge against the option price derivative $`\frac{^3f}{S^2\alpha _i^j}`$, which is more than we need. Because the derivative $`\frac{}{S}`$ has no unique discrete representation on the trinomial tree, the derivative $`\frac{^2f}{S\alpha _i^j}`$ also has no unique representation. However, we need not dwell too much on this technicality, as we will see that even our potentially too undemanding set of Eqs. (0.8) can not be satisfied in a three-step world and beyond. If we want to be hedged against local volatility movements only to first order, the constraints are $`\mathrm{\Lambda }^{}\left(\mathrm{\Pi }(m)\{1\}\right)=\mathrm{\Lambda }^{}\left(\mathrm{\Pi }(m)\{0\}\right)=\mathrm{\Lambda }^{}\left(\mathrm{\Pi }(m)\{1\}\right)`$ $`{\displaystyle \frac{\mathrm{\Lambda }^{}\left(\mathrm{\Pi }(m)\{0\}\right)}{\alpha _i^j}}=0`$ $`i=m+1,\mathrm{},n1,j=\mathrm{\Pi }_m(m)(im),\mathrm{},\mathrm{\Pi }_m(m)+(im)`$ (0.9) We will show that the constraints Eqs. (0.9) can always be satisfied. The algebra of the problem is very involved, and from the four-step or even three-step world upwards the equations are far too many to analyse all possible tractor hedges without the help of an algebraic calculations program. An example Mathematica<sup>TM</sup> notebook can be requested from the author. ### 5.5 Example: A two-step world As a first example, we consider a two-step tree where we are short $`K`$ tractor options $`f_{\{1,1\}}`$. Later, we shall also consider the most general exotic option in this two-step trinomial world. The zero-time hedged portfolio is $$\mathrm{\Lambda }(\mathrm{})=f(\mathrm{})+\omega (\mathrm{})S_0+w_1^1(\mathrm{})A_1^1(\mathrm{})\text{ }$$ $$+w_2^2(\mathrm{})A_2^2(\mathrm{})+w_2^1(\mathrm{})A_2^1(\mathrm{})+w_2^0(\mathrm{})A_2^0(\mathrm{})$$ After one step, before re-hedging, we have $$\mathrm{\Lambda }(\mathrm{\Pi }(1))=f(\mathrm{\Pi }(1))+\omega (\mathrm{})S_1^{\mathrm{\Pi }_1(1)}+w_1^1(\mathrm{})A_1^1(\mathrm{\Pi }(1))\text{ }$$ $$\text{ }+w_2^2(\mathrm{})A_2^2(\mathrm{\Pi }(1))+w_2^1(\mathrm{})A_2^1(\mathrm{\Pi }(1))+w_2^0(\mathrm{})A_2^0(\mathrm{\Pi }(1))$$ In full $`\mathrm{\Lambda }(\{1\})`$ $`=`$ $`K{\displaystyle \frac{\alpha _1^1(1)}{1+e^u}}+\omega (\mathrm{})S_0e^u+w_1^1(\mathrm{})`$ $`+w_2^2(\mathrm{}){\displaystyle \frac{\alpha _1^1(1)}{1+e^u}}+w_2^1(\mathrm{})(1\alpha _1^1(1))+w_2^0(\mathrm{}){\displaystyle \frac{\alpha _1^1(1)}{1+e^u}}`$ $`\mathrm{\Lambda }(\{0\})`$ $`=`$ $`\omega (\mathrm{})S_0+w_2^1(\mathrm{}){\displaystyle \frac{\alpha _1^0(1)}{1+e^u}}+w_2^0(\mathrm{})(1\alpha _1^0(1))`$ $`\mathrm{\Lambda }(\{1\})`$ $`=`$ $`\omega (\mathrm{})S_0e^u+w_2^0(\mathrm{}){\displaystyle \frac{\alpha _1^1(1)}{1+e^u}}`$ If we attempt to hedge to second order, we solve in all the Eqs. (0.8). Each of the portfolios depends only on one of the future local volatilities, so to first order all the first derivatives with respect to the local volatilities have to be zero, which leads to $$w_2^1(\mathrm{})=w_2^0(\mathrm{})=0,w_2^2(\mathrm{})=Ke^u$$ After that, the condition $`\mathrm{\Lambda }^{}(\{0\})=\mathrm{\Lambda }^{}(\{1\})`$ leads to $`\omega (\mathrm{})=0`$, and $`\mathrm{\Lambda }^{}(\{0\})=\mathrm{\Lambda }^{}(\{1\})`$ to $`w_1^1=0`$. So the only hedge we put on initially is $`e^uK`$ units of $`A_2^2`$. We also see that this makes the portfolio worthless initially. If the first step of the stock-price path is flat or down, the portfolio is patently worth-less. If it is up, the portfolio has also value zero, regardless of the actual value of the remaining local volatility in the future cone, $`\alpha _1^1`$. The hedge portfolio for this step is then trivially calculated by $`\mathrm{\Lambda }^{}(\{1,1\})=\mathrm{\Lambda }^{}(\{1,0\})=\mathrm{\Lambda }^{}(\{1,1\})`$. The solution is $`\omega (\{1\})=\frac{K}{S_0(e^u1)},w_2^2(\{1\})=e^uK`$. So the entire hedging strategy was found easily. Note that the hedge constraints were all linear in the local volatilities, because the constraints are formulated at the last time-slice. Any exotic option will have collapsed to an Arrow Debreu option by that time. Thus, the $`\sigma `$-algebra generated by the sets $`𝒟[0,0,1,j]𝒫\left[\mathrm{\Pi }(m)\right]`$ and the $`\sigma `$-algebra generated by the paths still relevant after the first time-step (i.e. when the hedging conditions need to be matched) are identical. Thus we expect the two-step world to provide a hedging strategy for all tractors. To demonstrate this, we can work out the whole hedged trading strategy. Let us write the general two-step exotic as a linear combination $$f=\underset{i,j=1}{\overset{1}{}}K_{\{i,j\}}𝒯_{\{i,j\}}.$$ Then the part of the hedging strategy relevant to the first time-step is $`\omega (\mathrm{})`$ $`=`$ $`{\displaystyle \frac{K_{\{0,0\}}K_{\{1,1\}}+e^uK_{\{1,0\}}}{(1e^u)S_0}}+{\displaystyle \frac{e^{2u}K_{\{1,1\}}}{(1e^u)S_0}}`$ $`w_1^1(\mathrm{})`$ $`=`$ $`K_{\{1,0\}}K_{\{0,1\}}e^uK_{\{0,1\}}+e^uK_{\{1,0\}}`$ $`w_2^0(\mathrm{})`$ $`=`$ $`K_{\{1,1\}}(1+e^u)K_{\{1,0\}}+e^uK_{\{1,1\}}`$ $`w_2^1(\mathrm{})`$ $`=`$ $`K_{\{0,1\}}(1+e^u)K_{\{0,0\}}+e^uK_{\{0,1\}}+(1+e^u)w_2^0`$ $`w_2^2(\mathrm{})`$ $`=`$ $`K_{\{1,1\}}(1+e^u)K_{\{1,0\}}+e^uK_{\{1,1\}}+(1+e^u)w_2^1e^uw_2^0`$ It is easy to check that with these parameters, the value of the portfolio is indeed independent of all the future volatilities. In all the cases the final hedge is trivial, and we will not demonstrate how to compute it. The reason that the hedge is trivial on the last time-step is that all the tractor options will have collapsed to Arrow-Debreu options. For this reason the problem of finding an exactly replicating portfolio in the two-step world is solvable. Note that a three-step world, however, collapses to a two-step world after the first step, and one third of all the tractor options become worthless at that time. The remaining ones are still two-step tractors. We will demonstrate by the example of a particular tractor option that the general three-step exotic can not be replicated to second order in the stock-price and local volatilities with a dynamic trading strategy of vanilla products. ### 5.6 Hedging Three-Step Tractors Imagine we want to hedge $`K`$ times the tractor $`𝒯_{\{0,0,0\}}`$ in a three-step world, i.e. the exotic which pays off $`K`$ units only if the stock never moves at all. For convenience, we will generally omit the arguments of $`w`$, and $`\alpha `$. Then the portfolios after one step are $`\mathrm{\Lambda }^{}(\{1\})`$ $`=`$ $`w_1^1+{\displaystyle \frac{\alpha _1^1}{1+e^u}}w_2^0+(1\alpha _1^1)w_2^1+{\displaystyle \frac{\alpha _1^1}{1+e^u}}w_2^2+{\displaystyle \frac{\alpha _1^1\alpha _2^0}{(1+e^u)^2}}w_3^1`$ $`+{\displaystyle \frac{\alpha _1^1(1\alpha _2^0)+(1\alpha _1^1)\alpha _2^1}{1+e^u}}w_3^0`$ $`+\left({\displaystyle \frac{\alpha _1^1(\alpha _2^2+\alpha _2^0)}{(1+e^u)(1+e^u)}}+(1\alpha _1^1)(1\alpha _2^1)\right)w_3^1`$ $`+{\displaystyle \frac{\alpha _1^1(1\alpha _2^2)+(1\alpha _1^1)\alpha _2^1}{1+e^u}}w_3^2+{\displaystyle \frac{\alpha _1^1\alpha _2^2}{(1+e^u)^2}}w_3^3+e^u\omega S_0`$ $`\mathrm{\Lambda }^{}(\{0\})`$ $`=`$ $`(1\alpha _1^0)w_2^0+{\displaystyle \frac{\alpha _1^0}{1+e^u}}w_2^1+{\displaystyle \frac{\alpha _1^0(1\alpha _2^1)+(1\alpha _1^0)\alpha _2^0}{1+e^u}}w_3^1`$ $`+\left({\displaystyle \frac{\alpha _1^0(\alpha _2^1+\alpha _2^1)}{(1+e^u)(1+e^u)}}+(1\alpha _1^0)(1\alpha _2^0)\right)w_3^0`$ $`+{\displaystyle \frac{\alpha _1^0(1\alpha _2^1)+(1\alpha _1^0)\alpha _2^0}{1+e^u}}w_3^1+{\displaystyle \frac{\alpha _1^0\alpha _2^1}{(1+e^u)^2}}w_3^2`$ $`+\omega S_0K(1\alpha _1^0)(1\alpha _2^0)`$ $`\mathrm{\Lambda }^{}(\{1\})`$ $`=`$ $`{\displaystyle \frac{\alpha _1^1}{1+e^u}}w_2^0+\left({\displaystyle \frac{\alpha _1^1(\alpha _2^0+\alpha _2^2)}{(1+e^u)(1+e^u)}}+(1\alpha _1^1)(1\alpha _2^1)\right)w_3^1`$ $`+{\displaystyle \frac{\alpha _1^1(1\alpha _2^0)+(1\alpha _1^1)\alpha _2^1}{1+e^u}}w_3^0+{\displaystyle \frac{\alpha _1^1\alpha _2^0}{(1+e^u)^2}}w_3^1+e^u\omega S_0`$ Let us try to find hedge ratios so as to satisfy the Eqs. (0.8). We will not illustrate all the tedious algebra involved, but we present one possible way to proceed, giving only some intermediate steps in the solution. First consider $`\frac{\mathrm{\Lambda }^{}(\{0\})}{\alpha _2^1}=0`$, which leads to $`w_3^2=e^uw_3^0+(1+e^u)w_3^1`$. Next, $`\frac{\mathrm{\Lambda }^{}(\{0\})}{\alpha _2^0}=0`$, gives $`w_3^1=(1+e^u)Ke^uw_3^1+(1+e^u)w_3^0`$. $`\frac{\mathrm{\Lambda }^{}(\{0\})}{\alpha _2^1}=0`$ leads to $`w_3^0=(1+e^u)w_3^1`$. Next, $`\frac{\mathrm{\Lambda }^{}(\{1\})}{\alpha _2^2}=\frac{\mathrm{\Lambda }^{}(\{1\})}{\alpha _2^2}`$ yields $`w_3^1=0`$. Demanding that $`\frac{\mathrm{\Lambda }^{}(\{1\})}{\alpha _2^2}=\frac{\mathrm{\Lambda }^{}(\{1\})}{\alpha _2^2}`$ then gives $`w_3^3=(1+e^u+e^{2u})(1+e^u)K`$. The simplified solutions so far are thus $`w_3^0=w_3^1=0`$, $`w_3^1=(1+e^u)K`$, $`w_3^2=(1+e^u)^2K`$, and $`w_3^3=(1+e^u+e^{2u})(1+e^u)`$. This settles the hedge ratios for the longest-living Arrow-Debreu prices. Next, we satisfy $`\frac{\mathrm{\Lambda }^{}(\{1\})}{\alpha _1^1}=\frac{\mathrm{\Lambda }^{}(\{1\})}{\alpha _1^1}`$ by setting $`w_2^2=e^uw_2^0+(1+e^u)w_2^1e^u\alpha _2^0K`$. The fact that the hedge ratio is dependent on a local volatility is not really a problem. We need to be careful, however, not to treat any $`\alpha `$’s which appear in a hedge ratio as variable when calculating other hedge-ratios because the implicit assumption is always that, while the stock price or the volatility are allowed to change instantaneously, the hedge ratios must stay constant until after the information on the new levels of stock price and volatilities has become known. With this in mind, the condition $`\frac{\mathrm{\Lambda }^{}(\{1\})}{\alpha _1^1}=\frac{\mathrm{\Lambda }^{}(\{1\})}{\alpha _1^1}`$ leads to $`w_2^0=K\alpha _2^0`$. Now, without demanding to fulfil the remaining Eqs. (0.8), we are already lead to an inevitable violation of one part of Eqs. (0.8), namely that $`\frac{\mathrm{\Lambda }^{}(\{1\})}{\alpha _2^0}\frac{\mathrm{\Lambda }^{}(\{1\})}{\alpha _2^0}=\frac{K}{1+e^u}(\alpha _1^1e^u\alpha _1^1)`$. There is no further hedge ratio that we can adjust to remedy this problem <sup>11</sup><sup>11</sup>11 In fact, we can just stubbornly proceed to solve all the other conditions in Eqs. (0.8), namely $`\frac{\mathrm{\Lambda }^{}(\{1\})}{\alpha _1^0}=0`$ and $`\mathrm{\Lambda }^{}(\{1\})=\mathrm{\Lambda }^{}(\{0\})=\mathrm{\Lambda }^{}(\{1\})`$, which give the hedge ratios $`w_2^1=(1+e^u)\alpha _2^0K`$, $`w_2^2=(1+e^u)^2\alpha _2^0K`$, $`w_1^1=0`$, and $`\omega =\frac{1\alpha _2^0}{1e^u}K`$, but this will not change the nature of the problem.. Now we can of course do the same analysis for all the 27 tractor options in this small trinomial world, but we do this with Mathematica<sup>TM</sup>. It turns out that in general, all but one of the second-order conditions of the Eqs. (0.8) can be satisfied and hedging to first order is possible. ### 5.7 Analysis of Larger Trinomial Worlds When analysing larger trinomial worlds, we quickly run into constraints in terms of processing power. Our Mathematica<sup>TM</sup> implementation frequently runs out of memory and needs copious manual intervention when analysing a five-step trinomial world. Nevertheless, some valuable observations can be made even with our calculations limited to a five-step and smaller trinomial worlds. The number of violated constraints in Eqs. (0.8) increases very quickly in going from a three-step to a four-step and five-step world. Obviously, the constraints arising in an $`n`$-step world are a subset of the constraints in any larger trinomial worlds, so in general we can not hope to find an exact second order hedge in any larger trinomial world. The question is, however, whether these terms matter, or whether their number becomes less and less significant when compared to the relevant terms. Let us define the number of un-satisfiable constraints in Eqs. (0.8) for a particular tractor $`𝒯_\pi `$ as $`\nu _\pi `$. Then we can for example measure the number $$N_n=\underset{\pi 𝒫_n}{}\nu _\pi ,$$ where we have given the set $`𝒫`$ a subscript $`n`$ to indicate how many steps the trinomial world has. We find $$\{N_1,N_2,\mathrm{}\}=\{0,0,21,210,1344,\mathrm{}\}$$ Note that on a 300 MHz/192 MB personal computer, the last number took about 12 hours to compute in Mathematica<sup>TM</sup>. The estimated time for the next element in the series is 160 hours, needing constant human intervention. So we have to make our conclusion from this limited sequence. In fact this sequence already indicates what kind of severity of mishedge we might expect in the continuum limit: In an $`n`$-step tree, there are $`n^21`$ local volatilities in the future cone of the origin, which means there are $`n^2`$ stochastic parameters to consider. This gives $`\frac{n^2(n^2+1)}{2}`$ P&L terms in the continuous trading equation arising from the quadratic variations of the stochastic variables alone. Assuming that the total vega-risk of vanilla options should not be affected by the discretisation, the relative importance of each particular quadratic term that can not be hedged is thus weighed down by a factor $`1/(n^2(n^2+1))`$ on average, when we compare different trinomial worlds. There are $`3^n`$ tractors in each $`n`$-step trinomial world, so the relative importance of each tractor, when defining an exotic option, will be weighed down by a factor $`3^n`$. As a rule of thumb, the relevance of all unsatisfied hedge constraints stays constant if, for large $`i`$, $`N_i3^ii^2(i^2+1)`$. Starting with $`N_3=21`$, the corresponding estimate of $`N_4`$ would be $`3N_3\frac{4^2(4^2+1)}{3^2(3^2+1)}=190.4`$, which is somewhat below the actual $`N_4`$, while the estimate of $`N_5`$ would be $`9N_3\frac{5^2(5^2+1)}{3^2(3^2+1)}=1362`$, which slightly larger than the actual $`N_5`$. With the limited sequence we were able to compute, it thus seems very plausible that the relative importance of the un-hedgeable terms does not disappear in the continuum limit. Lastly, we note that it is of course possible to construct ‘exotics’, i.e. linear combination of tractors, which do not exhibit any quadratic or linear volatility exposure after an optimal hedge is put in place. A trivial example are the European vanilla options themselves; another, less trivial, example are the tractors $`𝒯_{\{1,1,1,\mathrm{},1,1,i\}}i\{1,0,1\}`$ and $`𝒯_{\{1,1,1,\mathrm{},1,1,i\}}i\{1,0,1\}`$, and linear combinations thereof. Also, arbitrary tractors may often have more complicated linear combinations which are insensitive to local volatilities to second order. In the set of all exotic options, however, these will form a set of measure zero: The linear vector space spanned by the set of all tractor options is the space of all contingent claims on the underlying asset. The space spanned by all independent linear combinations which have no volatility exposure (to second order) after hedging with vanilla options is a subspace of that set. If any tractor option at all has un-hedgeable local-vega risk, this subspace is a true subspace, and thus forms a set of measure zero in the space of all contingent claims. Thus we have shown that – out of all conceivable contingent claims – almost all are unhedgeable with a trading strategy using vanilla options only. The severity of the problem is not mitigated in the continuum limit. ## 6 Discussion and Conclusions We have demonstrated that, in an arbitrage-free market, pricing exotic options with a DLVF model (implied tree) is inconsistent with the assumption of unknown future volatilities, inasmuch as the LVF implied in the European vanilla options market does not fully compensate for irreplicable volatility risks exhibited by exotics if vanilla options are the only instruments used for hedging volatility exposure. Intuitively, we can interpret this result in the following manner: Local volatility gadgets exist, but they do not rid us of second-order local volatility exposure, which generates steady additional (positive or negative) P&L over time. This P&L is not priced in <sup>12</sup><sup>12</sup>12 Furthermore, the implied tree generally gives no description of how the best possible hedge can be attained even within these limitations. This, however, is a purely practical issue.. By contrast, a market where the volatility process has only linear variation does not give us the second-order P&L terms, but is not arbitrage-free. A market where the volatility has no variation at all gives no vega-P&L problems at all, and is arbitrage free. It is under the assumption of this market that the DLVF model is constructed, which is theoretically self-consistent, but not consistent with reality . We stress that the lack of a second-order vega-hedge is not model-specific: even a stochastic volatility model would create that, as long as the volatilities which we try to imply are not substantially fewer than the number of vanilla option prices one calibrates the model to. However, in stochastic volatility models and jump-diffusion models the onus of pricing in volatility risk is not on the implied volatilities alone. Instead, the price of such risks can be calibrated to the market and particular model parameters exist to absorb this calibration explicitly. Low-parametric SLVF models do therefore not require the volatility risk to be effectively integrated out and – in some mysterious way – absorbed into the DLVF calibration. We have only considered LVF models. It should be noted, however, that what we pointed out here is a general concern which needs to be addressed for any asset-price model one is willing to consider: The model must price exotics consistently with the market parameters that one is calibrating the model to. As we have seen here, it is not always easy to convince oneself whether or not this demand is actually met by a proposed model. All we have discovered in this paper is that it is not possible to satisfy this condition when using an LVF model calibrated to vanilla option prices if (a) we only use the asset and vanilla options as hedge instruments, (b) volatilities are not completely deterministic, (c) the market is arbitrage-free, and (d) the number of local volatilities is of the order of the number of vanilla options available for model calibration. As this is a central conclusion, we sum it up as another ###### Theorem 3 Vanilla options do not complete the market in the most general SLVF model. However, it is possible that no single model price process satisfies this demand if we use one and the same calibration recipe regardless of what contingent claim we want to price. Thus, we must find a proof for a different model that it does price exotics in line with the vanilla options market. Possible resolutions of this dilemma might be * A low-parametric LVF surface calibration which includes volatility risk in what must essentially be an SLVF model. * A different calibration using quoted exotic options as well as vanilla European options. In this case we must also allow the use of these exotics as hedge instruments. * A jump-diffusion model, or a combination of jump-diffusion with SLVF or with DLVF as in ref. * A model which makes no self-consistent assumptions about the underlying process at all, but proves in other ways that it prices (maybe only specific) exotics ‘in line with’ the market. This amounts to finding a model-independent hedging strategy <sup>13</sup><sup>13</sup>13 Since one makes no assumptions about the underlying process, this model can not be general enough to allow us to price all conceivable exotics. As an example of the concept just note that it is trivial to price European options with arbitrary payoffs from vanilla European options in a model-independent way.. There is certainly no guarantee that any of the first three resolutions would work for all possible exotics, and the exit route proposed in the last item is likely to involve substantially different derivations for each different class of exotic options. Also, there is no guarantee that a model-independent hedging-solution exists for any particular exotic option. Faced with the task of pricing a particular exotic option while not having a bullet-proof hedging model, it must be of paramount concern to know the un-priced risks, whatever the model. The DLVF model gives us no information whatsoever on this. However, ideas such as the ‘uncertain volatility models’ of Avellaneda, Levy, and Parás, and of Lyons have emerged in an attempt to avoid the perils of trusting a single underlying-process to price all exotics correctly . The main premise of their approach is that the exotic pricing should be performed as model-independently as possible, by using optimum amounts of standard traded options to statically hedge out as much of the payoff as possible; thereafter the residual is delta-hedged in order to obtain the tightest possible worst/best price spread for the option within the assumption of a ‘certainty interval’ or confidence interval for future volatilities . In summary, we should never trust a model-process of the underlying to price a particular exotic correctly, unless we can prove this to be the case <sup>14</sup><sup>14</sup>14 The best way to prove that the model works is by analysing the hedging strategy that the model would suggest. While we saw in this paper that it is difficult to extract a hedging strategy from an DLVF model, other models may lend themselves more naturally to an interpretation in terms of the proposed hedging method.. The way in which the DLVF model is often trusted to even price all conceivable exotic options correctly – while we have shown that the set of correctly priced options is infinitely smaller (of measure zero in the space of all exotics) – should serve a warning. The effective theory works for almost none of the conceivable derivatives. As a by-product of the proof, we have also found that any tradable asset, if it is stochastic, must have at least quadratic variation – and will therefore produce a $`\mathrm{\Gamma }`$-like P&L on instruments with non-linear exposure to that asset – if the market is arbitrage-free. This result complements the observations of the HJM model and of Derman and Kani that the quadratic variation of their particular diffusive models determines the drift completely if the condition of no arbitrage is imposed . Even in the context of jump-diffusion models as for example in , the drift is required to be deterministic for the same reason. Since the drift cannot be stochastic in the absence of arbitrage, we must either have at least quadratic variation or full determinism.
warning/0001/hep-th0001149.html
ar5iv
text
# ON ACTION FUNCTIONALS FOR INTERACTING BRANE SYSTEMS ## Introduction Recently an approach to obtain a supersymmetric action functional for a coupled system of interacting superbranes has been proposed . The aim of the study was to construct a basis for a future search for new solitonic solutions of worldvolume equations of superbranes and of supergravity equations with a complicated system of brane sources. The system of open fundamental superstring ending on a super–Dp-brane was considered in as a generic case of interacting brane system in the frame of the ’brane democracy’ concept . On the other hand, the manifestly supersymmetric description of such system can be useful in a search for a quasiclassical effective action of the system of coincident branes which are conjectured to carry non-Abelian gauge fields (non-Abelian Dirac–Born–Infeld action ). In this contribution we would like to discuss the key issues of our approach by studying a pure bosonic limit of the interacting brane system. We elaborate the case of open bosonic string with endpoints living on the dynamical bosonic Dp–brane and discuss the relation of the approach with the standard description where the Dp-branes are treated as rigid $`p`$-dimensional hyperplanes. ## 1 Standard description of open string To clarify the problem and to establish the notations let us begin with the standard consideration of the open bosonic string in flat D–dimensional space–time $`\underset{¯}{}^D`$. The action can be written as $$S_{str}=\frac{1}{2}_^2d^2\xi \sqrt{g}g^{mn}(\xi )_m\widehat{X}^{\underset{¯}{m}}(\xi )_n\widehat{X}^{\underset{¯}{n}}(\xi )\eta _{\underset{¯}{m}\underset{¯}{n}}.$$ (1) Here $`\{(\xi ^m)\}=\{(\tau ,\sigma )\}`$ ($`m=0,1`$) are local coordinates on the string worldsheet $`^2`$, $`\widehat{X}^{\underset{¯}{m}}(\xi )`$ ( $`\underset{¯}{m}=0,\mathrm{},(D1)`$) are coordinate (embedding) functions which define (locally) an embedding of the worldsheet $`^2`$ into the D–dimensional space–time (target) $`\underset{¯}{}^D`$ $$X^{\underset{¯}{m}}=\widehat{X}^{\underset{¯}{m}}(\xi ):^2\underset{¯}{}^D.$$ (2) The flat metric of $`\underset{¯}{}^D`$ is chosen to be ’mostly minus’ $`\eta _{\underset{¯}{m}\underset{¯}{n}}=diag(+,,\mathrm{},)`$. When the worldsheet has a boundary $`^2`$, the variation of the action (1) acquires the form ($`d\xi ^md\xi ^n=\epsilon ^{mn}d^2\xi `$, $`(d\xi )^2=1/2\epsilon _{mn}d\xi ^md\xi ^n=d^2\xi `$) $$\delta S_{str}=\frac{1}{2}_^2d^2\xi \sqrt{g}\delta g^{mn}(\xi )\left(_m\widehat{X}^{\underset{¯}{m}}_n\widehat{X}_{\underset{¯}{m}}\frac{1}{2}g_{mn}g^{kl}_k\widehat{X}^{\underset{¯}{m}}_l\widehat{X}_{\underset{¯}{m}}\right)$$ $$_^2d^2\xi \delta \widehat{X}_{\underset{¯}{m}}_m\left(\sqrt{g}g^{mn}_n\widehat{X}^{\underset{¯}{m}}\right)_^2𝑑\xi ^m\epsilon _{mn}\sqrt{g}g^{nk}_k\widehat{X}^{\underset{¯}{m}}\delta \widehat{X}_{\underset{¯}{m}}.$$ (3) The variation with respect to the auxiliary intrinsic metric $`g_{mn}(\xi )`$ results in the equation $$_m\widehat{X}^{\underset{¯}{m}}(\xi )_n\widehat{X}_{\underset{¯}{n}}(\xi )=\frac{1}{2}g_{mn}(\xi )g^{pq}(\xi )_p\widehat{X}^{\underset{¯}{m}}(\xi )_q\widehat{X}_{\underset{¯}{n}}(\xi ).$$ (4) One easily finds that the trace part of Eq. (4) is satisfied identically. This is the Noether identity reflecting the gauge Weyl symmetry of the action (1) $$g_{mn}(\xi )g_{mn}^{}(\xi )=e^\mathrm{\Lambda }g_{mn}(\xi ),X^{\underset{¯}{m}}(\xi )X^{\underset{¯}{m}}(\xi ).$$ (5) With this symmetry one can fix the gauge $$g^{mn}(\xi )_m\widehat{X}^{\underset{¯}{m}}(\xi )_m\widehat{X}_{\underset{¯}{n}}(\xi )=2.$$ (6) Then Eq. (4) becomes $$g_{mn}(\xi )=_m\widehat{X}^{\underset{¯}{m}}(\xi )_m\widehat{X}_{\underset{¯}{n}}(\xi ),$$ (7) and implies that the metric is induced by the embedding. The usual way to deal with the coordinate variation of the action (1) (presented in the second line of Eq. (3)) is to assume that the bulk and the boundary inputs should vanish separately. Then one arrives at the well known result that * proper equations of motion have the form $$_m\left(\sqrt{g}g^{mn}(\xi )_n\widehat{X}^{\underset{¯}{m}}(\xi )\right)=0;$$ (8) * the boundary conditions should be chosen in a way which provides $$d\xi ^m\epsilon _{mn}\sqrt{g}g^{nk}_k\widehat{X}^{\underset{¯}{m}}\delta \widehat{X}_{\underset{¯}{m}}|_{^2:\xi ^m=\stackrel{~}{\xi }^m(\tau )}=0$$ (9) where the functions $`\stackrel{~}{\xi }^m(\tau )`$ define parametrically the ’embedding’ of (a connected piece of) the world sheet boundary $`^2`$ into the worldsheet $$\xi ^m=\stackrel{~}{\xi }^m(\tau ):^2^2.$$ (10) To satisfy Eq. (9) in a simple way one can assume that for some part $`\widehat{X}^i(\xi )`$ of the embedding functions $$\widehat{X}^{\underset{¯}{m}}=(\widehat{X}^a,\widehat{X}^i),a=0,\mathrm{}p,i=(p+1),\mathrm{},(D1)$$ (11) the Neumann boundary conditions are imposed $$d\xi ^m\epsilon _{mn}\sqrt{g}g^{nk}_k\widehat{X}^a|_{\xi ^m=\stackrel{~}{\xi }^m(\tau )}=0,_{}\widehat{X}^a|_^2=0,$$ (12) while the remaining equations from the set (9) are satisfied due to the Dirichlet boundary conditions. The latter can be treated as ones imposed on the variation $`\delta \widehat{X}^i(\stackrel{~}{\xi }^m(\tau ))=0`$ or, equivalently, as the boundary conditions imposed on the coordinate functions $`\widehat{X}^i(\xi )`$ directly $$\widehat{X}^i(\stackrel{~}{\xi }^m(\tau ))=\stackrel{~}{X}^i(=const)\delta \widehat{X}^i(\stackrel{~}{\xi }^m(\tau ))=0$$ (13) Note that for the Dirichlet boundary problem the equations (8) appear formally only on the proper open subset of the worldsheet, i.e. only outside the boundary, because the variation on the boundary vanishes just due to Eq. (13). In the Neumann problem we can regard Eq. (8) as valid on the whole worldsheet. But in this case we shall assume that the boundary conditions (12) are imposed after the variation of the action has been performed (as we did not restrict the variations by the condition (12) when derived the equations of motion). The boundary conditions (12), (13) certainly break Lorentz invariance for $`p(D1)`$. Such a breaking reflects the existence of some $`d=(p+1)`$-dimensional defects where the string endpoints move. These are worldvolumes of Dp-branes considered as flat hyperplanes. ## 2 Extended variational problem However, we can proceed in a different manner which we will call extended variational problem or extended variational approach. Let us introduce the current density distribution with support on the boundary of the worldsheet ($`d\xi ^md\xi ^n=d^2\xi ϵ^{mn}`$) $$j_1=d\xi ^m\epsilon _{mn}_^2𝑑\stackrel{~}{\xi }^n(\tau )\delta ^2\left(\xi \stackrel{~}{\xi }(\tau )\right)=\epsilon _{mn}d\xi ^mj^n,$$ (14) Then one easily find that $$_^2j_1\widehat{𝒜}_1=_^2\stackrel{~}{\widehat{𝒜}}_1$$ (15) holds for any one-form $`\widehat{𝒜}_1=d\xi ^m𝒜_m(\xi )`$ defined on the worldsheet, $`\stackrel{~}{\widehat{𝒜}}_1=d\stackrel{~}{\xi }^m(\tau )𝒜_m(\stackrel{~}{\xi }(\tau ))`$. With the use of $`j_1`$ the coordinate variation of the string action can be written in the form $$\delta S_{str}=_^2\left(d^2\xi _m\left(\sqrt{g}g^{mn}_n\widehat{X}^{\underset{¯}{m}}\right)j_1d\xi ^m\epsilon _{mn}\sqrt{g}g^{nk}_k\widehat{X}^{\underset{¯}{m}}\right)\delta \widehat{X}_{\underset{¯}{m}}$$ (16) The equations of motion are (7) and $$d^2\xi _m\left(\sqrt{g}g^{mn}(\xi )_n\widehat{X}^{\underset{¯}{m}}(\xi )\right)=j_1d\xi ^m\epsilon _{mn}\sqrt{g}g^{nk}_k\widehat{X}^{\underset{¯}{m}}$$ (17) $$_m\left(\sqrt{g}g^{mn}_n\widehat{X}^{\underset{¯}{m}}\right)=_^2𝑑\stackrel{~}{\xi }^m\epsilon _{mn}\sqrt{g}g^{nk}_k\widehat{X}^{\underset{¯}{m}}\delta ^2\left(\xi \stackrel{~}{\xi }(\tau )\right).$$ Thus no boundary conditions appear, but the equations of motion acquire a source localized at the boundary of the worldsheet. The result (17) of such an ’extended variational problem’ looks different from the standard one (8), (12), (13). To clarify the situation let us, for the moment, fix the conformal gauge and accept the parametrization where the string endpoints correspond to the values $`\sigma =0`$ and $`\sigma =\pi `$. Then the standard approach produces the free (Laplace) equations with boundary conditions $$(_\tau ^2_\sigma ^2)\widehat{X}^a(\tau ,\sigma )=0,_\sigma \widehat{X}^a|_^2=0,$$ (18) $$(_\tau ^2_\sigma ^2)\widehat{X}^i(\tau ,\sigma )=0,\widehat{X}^i(\tau ,0)=\stackrel{~}{X}_0^i,\widehat{X}^i(\tau ,\pi )=\stackrel{~}{X}_\pi ^i,$$ (19) while the extended variational problem results in the equations with sources for both types of the coordinate functions $$(_\tau ^2_\sigma ^2)\widehat{X}^a(\tau ,\sigma )=_\sigma \widehat{X}^a\left(\delta (\sigma )\delta (\sigma \pi )\right)$$ (20) $$(_\tau ^2_\sigma ^2)\widehat{X}^i(\tau ,\sigma )=_\sigma \widehat{X}^i\left(\delta (\sigma )\delta (\sigma \pi )\right)$$ (21) It is evident that the problem (18) can be obtained as a particular case of (20) because, after imposing an additional condition $`_\sigma \widehat{X}^a|_^2=0`$, the r.h.s. of Eq. (20) vanishes. As in such a way we arrive at the free equations on the whole worldsheet, this corresponds just to the result of Neumann variation problem with the supposition that the boundary conditions are imposed after the variation has been performed (i.e. imposed ’by hand’). This does not hold for (19), because $`\widehat{X}^i|_^2=const`$ does not imply $`_\sigma \widehat{X}^i|_^2=0`$. Thus the value of Laplace operator acting on the coordinate function is not indefinite on the boundary, as it is in the classical Dirichlet problem, but is determined by the value of the derivative $`_\sigma \widehat{X}^i|_^2`$. The fact that the straightforward application of the extended variational approach does not produce the Dirichlet boundary problem even as a particular case does not look not so surprising if one remembers that it assumes that the variations of the coordinate functions are unrestricted everywhere, while the Dirichlet boundary conditions (13) come just as restrictions on the variations. ## 3 Lagrange multiplier approach for Dirichlet problem. If one would like to reproduce the Dirichlet problem (19) from the extended variational approach, one can try to incorporate the boundary conditions (13) with Lagrange multiplier one-form $`P_1^id\tau 𝒫^i(\tau )`$ into the action written in the conformal gauge $`g_{mn}=_mX^{\underset{¯}{m}}_nX_{\underset{¯}{m}}=\eta _{mn}diag(+1,1)`$ (cf. ) $$S_{str}=\frac{1}{2}_^2d^2\xi \left(_m\widehat{X}^a^m\widehat{X}_a_m\widehat{X}^i^m\widehat{X}^i\right)+_^2𝑑\tau 𝒫_i\left(\widehat{X}^i\left(\xi (\tau )\right)\stackrel{~}{X}^i\right)$$ (22) Then the Dirichlet boundary conditions are produced by the variation with respect to Lagrange multiplier $`𝒫^i(\tau )`$ or, more precisely, two Lagrange multipliers $`𝒫_i^0(\tau )`$ and $`𝒫_i^\pi (\tau )`$ (each living on the worldline of the corresponding endpoint of the string), while the dynamical equations for the coordinate functions $`\widehat{X}^i`$ read $$(_\tau ^2_\sigma ^2)\widehat{X}^i(\tau ,\sigma )=\left(_\sigma \widehat{X}^i𝒫_i^0\right)\delta (\sigma )\left(_\sigma \widehat{X}^i𝒫_i^\pi \right)\delta (\sigma \pi ).$$ (23) What one can see from the very beginning is the arbitrariness in the Lagrangian multipliers $`𝒫_i^{0,\pi }(\tau )`$ which cannot be fixed by any gauge symmetry of the action. So one can conclude that the Lagrange multipliers carry some degrees of freedom and some doubts may arise concerning the applicability of the Lagrangian multiplier method to the problem. However if one notes that i) Eq. (23) becomes free when considered outside the boundary (i.e. on the open proper subset of the string worldsheet), ii) the boundary conditions (13) are produced as equations of motion for Lagrange multipliers, iii) the action of the Laplace operator on the coordinate functions $`X^i`$ is indefinite on the boundary just due to the arbitrariness of the Lagrange multiplier, one easily concludes that the Dirichlet boundary problem is indeed reproduced. When one can find a solution for it (as it can be done in the case of string), then, substituting it into the equations, one can fix the value of the Lagrange multiplier $`𝒫^i(\tau )`$. This way to reproduce the Dirichlet boundary problem could be regarded as an artificial one. However it just provides the possibility to describe the system of open (super)string and dynamical (super-)Dp–brane and, more generally, of open (super–)brane(s) ending on closed ’host’ (super–)brane(s) on the level of (quasi)classical action functional . ## 4 Dynamical string with the ends on dynamical D-brane The action for free Dp-brane has the form $$S_{Dp}=_{^{p+1}}d^{p+1}\zeta \sqrt{|G|},|G|=()^pdet\left(G_{\stackrel{~}{m}\stackrel{~}{n}}\right),$$ (24) where $$G_{\stackrel{~}{m}\stackrel{~}{n}}_{\stackrel{~}{m}}\stackrel{~}{X}^{\underset{¯}{m}}(\zeta )_{\stackrel{~}{n}}\stackrel{~}{X}_{\underset{¯}{m}}(\zeta )_{\stackrel{~}{m}\stackrel{~}{n}}$$ (25) is the non-symmetric ’open string induced metric’ , the coordinate functions $`\widehat{X}^{\underset{¯}{m}}(\zeta ^{\stackrel{~}{m}})`$ define (locally) an embedding of the $`d=(p+1)`$-dimensional worldvolume $`^{p+1}=\{(\zeta ^{\stackrel{~}{m}})\}`$ ($`\stackrel{~}{m}=0,\mathrm{},p`$) of the Dp-brane into the D(=10)–dimensional space–time $`\underset{¯}{}^D`$ $$X^{\underset{¯}{m}}=\stackrel{~}{X}^{\underset{¯}{m}}(\zeta ):^{p+1}\underset{¯}{}^D,$$ (26) $$dA=\frac{1}{2}d\zeta ^{\stackrel{~}{m}}d\zeta ^{\stackrel{~}{n}}_{\stackrel{~}{n}\stackrel{~}{m}}$$ (27) is the field strength of the worldvolume gauge field of the Dp–brane $`A=d\zeta ^{\stackrel{~}{m}}A_{\stackrel{~}{m}}(\zeta )`$. With this notation the variation of the free Dp-brane action (24) can be written in the form $$\delta S_{Dp}=+_{^{p+1}}d^{p+1}\zeta _{\stackrel{~}{m}}\left(\sqrt{|G|}G^{1[\stackrel{~}{m}\stackrel{~}{n}]}\right)\delta A_{\stackrel{~}{n}}(\zeta )$$ (28) $$_{^{p+1}}d^{p+1}\zeta _{\stackrel{~}{m}}\left(\sqrt{|G|}G^{1(\stackrel{~}{m}\stackrel{~}{n})}_{\stackrel{~}{n}}\stackrel{~}{X}^{\underset{¯}{m}}\right)\delta \stackrel{~}{X}_{\underset{¯}{m}},$$ where $$G^{1mn}=G^{1(mn)}+G^{1[mn]}$$ is the matrix inverse to the open string metrics (25). To describe the interacting system of the open string and the dynamical D-brane on the level of action functional one has to provide the counterpart of (13) with coordinate functions of D-brane instead of constants $`\stackrel{~}{X}^i`$. Namely we have to impose an identification $$\widehat{X}^{\underset{¯}{m}}\left(\stackrel{~}{\xi }(\tau )\right)=\stackrel{~}{X}^{\underset{¯}{m}}\left(\widehat{\zeta }(\tau )\right),$$ (29) where $`\zeta ^{\stackrel{~}{m}}=\widehat{\zeta }^{\stackrel{~}{m}}(\tau )`$ defines an embedding of the worldline(s) of string endpoints (i.e. the boundary of the worldsheet) into the worldvolume of the Dp-brane $$\zeta ^{\stackrel{~}{m}}=\widehat{\zeta }^{\stackrel{~}{m}}(\tau ):^2^{p+1}.$$ (30) The problem is that, when (29) is taken into account the variations $`\delta \widehat{X}^{\underset{¯}{m}}`$ and $`\delta \stackrel{~}{X}^{\underset{¯}{m}}|_{\zeta =\widehat{\zeta }(\tau )}`$ cannot be treated as independent on the intersection $`^{1+p}^2=^2`$. Indeed, varying (29) one finds $$\delta \widehat{X}^{\underset{¯}{m}}|_{\xi =\stackrel{~}{\xi }(\tau )}+\delta \stackrel{~}{\xi }^m(\tau )_m\widehat{X}^{\underset{¯}{m}}\left(\stackrel{~}{\xi }(\tau )\right)=\delta \stackrel{~}{X}^{\underset{¯}{m}}|_{\zeta =\widehat{\zeta }(\tau )}+\delta \widehat{\zeta }^{\stackrel{~}{m}}(\tau )_{\stackrel{~}{m}}\stackrel{~}{X}^{\underset{¯}{m}}\left(\widehat{\zeta }(\tau )\right).$$ (31) A method to implement (29) into the action principle was proposed in . It consists in imposing the identification (29) with Lagrange multipliers one-form $`P_1^{\underset{¯}{m}}=d\tau 𝒫^{\underset{¯}{m}}(\tau )`$ into the action (cf. with (22)). Thus the action for a coupled system of open string and dynamical Dp-brane can be written as $$S=S_{str}+S_{Dp}+S_{int}.$$ (32) Here $`S_{str}`$ is determined by (1) but with open worldsheet whose boundary is a set of two (or more!) worldlines which are embedded into the worldvolume of the Dp–brane (30). $`S_{Dp}`$ is determined by Eq. (24). The last term in (32) $$S_{int}=_^2\left(\widehat{A}+P_{1\underset{¯}{m}}\left(\widehat{X}^{\underset{¯}{m}}(\stackrel{~}{\xi }(\tau ))\stackrel{~}{X}^{\underset{¯}{m}}(\widehat{\zeta }(\tau ))\right)\right).$$ (33) describes, in particular, the interaction of the string endpoints with the worldvolume gauge field of the Dp–brane $`A=d\zeta ^{\stackrel{~}{m}}A_{\stackrel{~}{m}}(\zeta )`$ whose pull-back on the intersection is $`\widehat{A}=d\widehat{\zeta }^{\stackrel{~}{m}}(\tau )A_{\stackrel{~}{m}}(\widehat{\zeta }(\tau ))=d\tau _\tau \widehat{\zeta }^{\stackrel{~}{m}}A_{\stackrel{~}{m}}(\widehat{\zeta })`$. The variation of the interaction term is $$\delta S_{int}=_^2\delta P_{1\underset{¯}{m}}\left[\widehat{X}^{\underset{¯}{m}}(\stackrel{~}{\xi }(\tau ))\stackrel{~}{X}^{\underset{¯}{m}}(\widehat{\zeta }(\tau ))\right]+$$ (34) $$+_^2P_{1\underset{¯}{m}}\left[\delta \widehat{X}^{\underset{¯}{m}}(\stackrel{~}{\xi }(\tau ))\delta \stackrel{~}{X}^{\underset{¯}{m}}(\widehat{\zeta }(\tau ))\right]+_^2𝑑\widehat{\zeta }^{\stackrel{~}{n}}(\tau )\delta A_{\stackrel{~}{n}}\left(\widehat{\zeta }(\tau )\right)+$$ $$+_^2\delta \widehat{\zeta }^{\stackrel{~}{m}}(\tau )\left[i_{\stackrel{~}{m}}P_{1\underset{¯}{m}}_{\stackrel{~}{m}}\stackrel{~}{X}^{\underset{¯}{m}}\right]+$$ $$+_^2\delta \widehat{\xi }^m(\tau )\left[P_{1\underset{¯}{m}}_m\widehat{X}^{\underset{¯}{m}}d\xi ^p\epsilon _{pn}\sqrt{g}g^{nk}_k\widehat{X}^{\underset{¯}{m}}_m\widehat{X}_{\underset{¯}{m}}\right].$$ The third and forth lines include the variation with respect to coordinate functions $`\widehat{\zeta }^{\stackrel{~}{m}}(\tau )`$ and $`\stackrel{~}{\xi }^m(\tau )`$ defining the embedding of the string boundary worldlines in the Dp–brane worldvolume (30) and in the worldsheet (10). To vary the whole action it is convenient to introduce the following current density distribution form (see and refs. therein)<sup>1</sup><sup>1</sup>1 In our conventions $`d\zeta ^{\stackrel{~}{m}_1}\mathrm{}d\zeta ^{\stackrel{~}{m}_{p+1}}=`$ $`\epsilon ^{\stackrel{~}{m}_1\mathrm{}\stackrel{~}{m}_{p+1}}(d\zeta )^{(p+1)}`$ $`()^p\epsilon ^{\stackrel{~}{m}_1\mathrm{}\stackrel{~}{m}_{p+1}}d^{p+1}\zeta `$, $`d\zeta _{\stackrel{~}{m}_1}^p\epsilon _{\stackrel{~}{m}_1\stackrel{~}{m}_2\mathrm{}\stackrel{~}{m}_{p+1}}d\zeta ^{\stackrel{~}{m}_2}\mathrm{}d\zeta ^{\stackrel{~}{m}_{p+1}}`$, $`\mathrm{}`$ $$j_p=d\zeta _{\stackrel{~}{m}}^p_{^{p+1}}𝑑\widehat{\zeta }^{\stackrel{~}{m}}(\tau )\delta ^{p+1}\left(\zeta \widehat{\zeta }(\tau )\right)=()^pd\zeta _{\stackrel{~}{n}}^pj^{\stackrel{~}{n}}.$$ (35) As $`j_pd\zeta ^{\stackrel{~}{m}}=d^{p+1}\zeta j^{\stackrel{~}{m}}`$, it is easy to see that $$_{^{1+3}}j_3\stackrel{~}{𝒜}_1=_{^{1+1}}\widehat{𝒜}_1.$$ (36) With the use of (35) and the worldsheet current density distribution (14) one can write the variation of the action (32) in the form ($`G_{\stackrel{~}{m}\stackrel{~}{n}}_{\stackrel{~}{m}}\stackrel{~}{X}^{\underset{¯}{m}}(\zeta )_{\stackrel{~}{n}}\stackrel{~}{X}_{\underset{¯}{m}}(\zeta )_{\stackrel{~}{m}\stackrel{~}{n}}`$ (25), $`G^{1\stackrel{~}{m}\stackrel{~}{q}}G_{\stackrel{~}{q}\stackrel{~}{n}}=\delta _{\stackrel{~}{n}}^{\stackrel{~}{m}}`$) $$\delta S=\frac{1}{2}_^2d^2\xi \sqrt{g}\delta g^{mn}(\xi )\left(_m\widehat{X}^{\underset{¯}{m}}_n\widehat{X}_{\underset{¯}{m}}\frac{1}{2}g_{mn}g^{kl}_k\widehat{X}^{\underset{¯}{m}}_l\widehat{X}_{\underset{¯}{m}}\right)$$ (37) $$_^2\delta \widehat{X}_{\underset{¯}{m}}\left[d^2\xi _m\left(\sqrt{g}g^{mn}_n\widehat{X}^{\underset{¯}{m}}\right)j_1\left(P_1^{\underset{¯}{m}}+d\xi ^m\epsilon _{mn}\sqrt{g}g^{nk}_k\widehat{X}^{\underset{¯}{m}}\right)\right]$$ $$+\underset{^2}{}\delta P_{1\underset{¯}{m}}\left[\widehat{X}^{\underset{¯}{m}}(\stackrel{~}{\xi }(\tau ))\stackrel{~}{X}^{\underset{¯}{m}}(\widehat{\zeta }(\tau ))\right]$$ $$\underset{^{p+1}}{}d^{p+1}\zeta \left[_{\stackrel{~}{m}}\left(\sqrt{|G|}G^{1[\stackrel{~}{m}\stackrel{~}{n}]}\right)j^{\stackrel{~}{n}}\right]\delta A_{\stackrel{~}{n}}(\zeta )$$ $$_{^{p+1}}\left[d^{p+1}\zeta _{\stackrel{~}{m}}\left(\sqrt{|G|}G^{1(\stackrel{~}{m}\stackrel{~}{n})}_{\stackrel{~}{n}}\stackrel{~}{X}^{\underset{¯}{m}}\right)+j_pP_1^{\underset{¯}{m}}\right]\delta \stackrel{~}{X}_{\underset{¯}{m}}.$$ Then the equations of motion become evident. They are * The Born–Infeld equation (with $`G_{mn}`$ defined by (25)) $$_m\left(\sqrt{|G|}G^{1[\stackrel{~}{m}\stackrel{~}{n}]}\right)=j^{\stackrel{~}{n}}_^2𝑑\widehat{\zeta }^{\stackrel{~}{n}}(\tau )\delta ^{p+1}\left(\zeta \widehat{\zeta }(\tau )\right).$$ (38) It includes the definite source localized on the intersection (and thus outside the intersection the equation is free). * Equations for the Dp-brane coordinate functions (with $`G_{mn}`$ defined by Eq. (25)) <sup>2</sup><sup>2</sup>2 To arrive at the expression for the r.h.s. one should perform some formal extension of the property (36), as the form $`P_1=d\tau 𝒫(\tau )`$ has not been defined as a pull–back of any 1-form on $`^{p+1}`$. Let us assume that such a form $`\stackrel{~}{P}_{1\underset{¯}{m}}=d\zeta ^{\stackrel{~}{m}}\stackrel{~}{𝒫}_{\underset{¯}{m}}`$ is introduced. Then, by definition, $`\stackrel{~}{P}_{1\underset{¯}{m}}(\widehat{\zeta }(\tau ))`$ $``$ $`d\tau _\tau \widehat{\zeta }^{\stackrel{~}{m}}\stackrel{~}{𝒫}_{\underset{¯}{m}}(\widehat{\zeta }(\tau ))=d\tau 𝒫_{\underset{¯}{m}}(\tau )`$ and one finds $`_{^{p+1}}j_p\stackrel{~}{P}_{1\underset{¯}{m}}=_^2d\tau 𝒫_{\underset{¯}{m}}(\tau ).\delta ^{p+1}(\zeta \widehat{\zeta }(\tau )`$. $$_m\left(\sqrt{|G|}G^{1(\stackrel{~}{m}\stackrel{~}{n})}_{\stackrel{~}{n}}\stackrel{~}{X}^{\underset{¯}{m}}\right)=_^2𝑑\tau 𝒫^{\underset{¯}{m}}(\tau )\delta ^{p+1}\left(\zeta \widehat{\zeta }(\tau )\right)$$ (39) * The equation for the string coordinate functions (with $`g_{mn}(\xi )`$ defined by (7) or, equivalently, by (4) and the Weyl symmetry (5)) $$d^2\xi _m\left(\sqrt{g}g^{mn}(\xi )_n\widehat{X}^{\underset{¯}{m}}(\xi )\right)=j_1\left(P_1^{\underset{¯}{m}}+d\xi ^m\epsilon _{mn}\sqrt{g}g^{nk}_k\widehat{X}^{\underset{¯}{m}}\right).$$ (40) ## 5 Reparametrization symmetry of the coupled system The variations of the action (32) with respect to the worldvolume and worldsheet embedding coordinate functions $`\delta \widehat{\zeta }(\tau )`$ (30) and $`\stackrel{~}{\xi }^m(\tau )`$ (10) is $$\delta _{\stackrel{~}{\xi },\widehat{\zeta }}S=_^2\delta \widehat{\zeta }^{\stackrel{~}{m}}(\tau )\left[i_{\stackrel{~}{m}}P_{1\underset{¯}{m}}_{\stackrel{~}{m}}\stackrel{~}{X}^{\underset{¯}{m}}\right]+$$ (41) $$+_^2\delta \stackrel{~}{\xi }^m(\tau )\left[P_{1\underset{¯}{m}}_m\widehat{X}^{\underset{¯}{m}}d\stackrel{~}{\xi }^p(\tau )\epsilon _{pn}\sqrt{g}g^{nk}_k\widehat{X}^{\underset{¯}{m}}_m\widehat{X}_{\underset{¯}{m}}\right].$$ The corresponding equations of motion $$i_{\stackrel{~}{m}}|_^2=P_{1\underset{¯}{m}}_{\stackrel{~}{m}}\stackrel{~}{X}^{\underset{¯}{m}}(\zeta (\tau ))$$ (42) $$d\widehat{\xi }^p(\tau )\epsilon _{pn}\sqrt{g}g^{nk}_k\widehat{X}^{\underset{¯}{m}}_m\widehat{X}_{\underset{¯}{m}}=P_{1\underset{¯}{m}}_m\widehat{X}^{\underset{¯}{m}}$$ (43) are dependent. To prove the dependence of Eq. (43) one should contract Eq. (40) with $`_l\widehat{X}_{\underset{¯}{m}}(\xi )`$ on the target space indices. Then, writing the left hand part as $$_m\left(\sqrt{g}g^{mn}(\xi )_n\widehat{X}^{\underset{¯}{m}}(\xi )\right)_l\widehat{X}^{\underset{¯}{m}}(\xi )=$$ (44) $$=_m\left(\sqrt{g}g^{mn}(\xi )_n\widehat{X}^{\underset{¯}{m}}(\xi )\right)_l\widehat{X}^{\underset{¯}{m}}(\xi ))(\sqrt{g}g^{mn}(\xi )_n\widehat{X}^{\underset{¯}{m}}(\xi ))_l_m\widehat{X}^{\underset{¯}{m}}(\xi )$$ and using Eq. (4) (which remains the same for the coupled system), one obtains (in arbitrary space-time dimension D) $$d^2\xi \sqrt{g}\frac{D2}{4}_l\left(\sqrt{g}g^{mn}(\xi )_m\widehat{X}^{\underset{¯}{m}}(\xi )_n\widehat{X}^{\underset{¯}{m}}(\xi )\right)=$$ (45) $$j_1\left(P_1^{\underset{¯}{m}}+d\xi ^m\epsilon _{mn}\sqrt{g}g^{nk}_k\widehat{X}^{\underset{¯}{m}}\right).$$ However, using the Weyl symmetry one can fix the gauge (6), where the l.h.s. of Eq. (45) vanishes, and we arrive at (43). To prove the dependence of Eq. (42) one should consider the longitudinal part of Eq. (39), i.e the result of the contraction of Eq. (39) with $`_{\stackrel{~}{m}}\stackrel{~}{X}^{\underset{¯}{m}}`$ on the target space indices $$_m\left(\sqrt{|G|}G^{1(\stackrel{~}{m}\stackrel{~}{n})}_{\stackrel{~}{n}}\stackrel{~}{X}^{\underset{¯}{m}}\right)_{\stackrel{~}{l}}\stackrel{~}{X}_{\underset{¯}{m}}=p^{\underset{¯}{m}}_{\stackrel{~}{l}}\stackrel{~}{X}_{\underset{¯}{m}}.$$ (46) Here we, for shortness, introduced the notation $`_^2𝑑\tau 𝒫^{\underset{¯}{m}}(\tau )\delta ^{p+1}\left(\zeta \widehat{\zeta }(\tau )\right)`$ $`=p^{\underset{¯}{m}}.`$ Dealing with the l.h.s. of Eq. (46) in the same way as in (44) and using the identities $$\delta _{\stackrel{~}{n}}^{\stackrel{~}{m}}=G^{1(\stackrel{~}{m}\stackrel{~}{k})}_{\stackrel{~}{k}}\stackrel{~}{X}^{\underset{¯}{m}}_{\stackrel{~}{n}}\stackrel{~}{X}_{\underset{¯}{m}}+G^{1[\stackrel{~}{m}\stackrel{~}{k}]}F_{\stackrel{~}{k}\stackrel{~}{n}},$$ $$G^{1[\stackrel{~}{m}\stackrel{~}{k}]}_{\stackrel{~}{k}}\stackrel{~}{X}^{\underset{¯}{m}}_{\stackrel{~}{n}}\stackrel{~}{X}_{\underset{¯}{m}}=G^{1(\stackrel{~}{m}\stackrel{~}{k})}F_{\stackrel{~}{k}\stackrel{~}{n}},$$ (which follow from $`G^{1\stackrel{~}{m}\stackrel{~}{k}}G_{\stackrel{~}{k}\stackrel{~}{n}}=\delta _{\stackrel{~}{n}}^{\stackrel{~}{m}}`$) one arrives at $$_m\left(\sqrt{|G|}G^{1(\stackrel{~}{m}\stackrel{~}{n})}_{\stackrel{~}{n}}\stackrel{~}{X}^{\underset{¯}{m}}\right)_{\stackrel{~}{l}}\stackrel{~}{X}_{\underset{¯}{m}}=_m\left(\sqrt{|G|}G^{1[\stackrel{~}{m}\stackrel{~}{k}]}\right)F_{\stackrel{~}{k}\stackrel{~}{n}}.$$ (47) Substituting Eqs. (46) and (38) we obtain from Eq. (47) $$_^2\left(i_{\stackrel{~}{m}}F_2+P_{1\underset{¯}{m}}_{\stackrel{~}{m}}\stackrel{~}{X}^{\underset{¯}{m}}(\zeta (\tau ))\right)\delta ^{p+1}\left(\zeta \widehat{\zeta }(\tau )\right)=0.$$ (48) Eq. (48) is equivalent to (42). The dependence of the equations (42), (43) should be regarded as Noether identities which reflect some gauge symmetries of the action. Reversing the arguments, we can consider (42), (43) as independent equations, but conclude that the longitudinal parts of Eqs. (39), (40) are dependent as they are in the case of the free string and Dp-brane. From such point of view we can state that the dependence of Eq. (43) and Eq. (42) reflects the reparametrization symmetries of the string worldsheet and the Dp-brane worldvolume which, hence, survive when coupling is ’switched on’. ## 6 Concluding remarks Thus, applying the method of Ref. we obtained the equations of motion (38) -(40), (29) for the coupled system of an open bosonic string and a dynamical bosonic Dp–brane where the string endpoints live. In such a way we illustrated the key issues of the approach and we are able to comment on its relation to the standard description of Dp–branes as rigid planes . The investigation of Eqs. (38) -(40), (29) and their supersymmetric generalizations will be the subject of the forthcoming paper. Note that such a study simplifies essentially when the Lorentz–harmonic formulations of the (super–)string and (super–)Dp–brane actions are used instead of (1), (24). ## Acknowledgments The authors are grateful to D. Sorokin and M. Tonin for interest in this work and many useful conversations. One of the authors (I.B.) thanks Padova Section of INFN for hospitality in Padova, where a part of this work has been done. A partial support from the INTAS Grant 96-308 and from the Ukrainian GKNT Grant 2.5.1/52 is acknowledged. ## References
warning/0001/cond-mat0001101.html
ar5iv
text
# Nonequilibrium Kinetic Ising Models: Phase Transitions and Universality Classes in One Dimension ## I Introduction The Ising model is a well known static, equilibrium model. Its dynamical generalizations, the kinetic Ising models, were originally intended to study relaxational processes near equilibrium states . Glauber introduced the single spin-flip kinetic Ising model, while Kawasaki constructed a spin-exchange version for studying the case of conserved magnetization. To check ideas of dynamic critical phenomena, kinetic Ising models were simple enough for producing analytical and numerical results, especially in one dimension, where also exact solutions could be obtained, and thus have proven to be a useful ground for testing ideas and theories . On the other hand, when attention turned to non-equilibrium processes, steady states, and phase transitions, kinetic Ising models were again near at hand. Nonequilibrium kinetic Ising models, in which the steady state is produced by kinetic processes in connection with heat baths at different temperatures, have been widely investigated , and results have shown that various phase transitions are possible under nonequilibrium conditions, even in one dimension (1d) (for a review see the article by Rácz in Ref. ). Also in the field of ordering kinetics, universality is strongly influenced by the conserved or non-conserved character of the order parameter. The scaling behaviour and characteristic exponents of domain growth are of central importance. Kinetic Ising models offer a useful laboratory again to explore the different factors which influence these properties . Most of these studies, however, have been concerned with the effects the nonequilibrium nature of the dynamics might exert on phase transitions driven by temperature. In some examples nonlocal dynamics can generate long-range effective interactions which lead to mean-field (MF)-type phase transitions (see, e.g., Ref. ). A different line of investigating nonequilibrium phase transitions has been via branching annihilating random walk (BARW) processes . Here particles chosen at random carry out random walks (with probability p) and annihilate pairwise on encounter. The increase of particles in ensured through production of $`n`$ offspring, with probability $`1p`$. Numerical studies have led to the conclusion that a general feature of BARW is a transition as a function of p between an active stationary state with non-zero particle density and an absorbing, inactive one in which all particles are extinct. The parity conservation of particles is decisive in determining the universality class of the phase transition. In the odd-$`n`$ case the universality class is that of directed percolation (DP), while the critical behavior is different for $`n`$ even . The first numerical investigation of this model was done by Takayasu and Tretyakov . This new universality class is often called parity conserving (PC) (however, other authors call it the DI or BAWe class ); we will also refer to it by this name. A coherent picture of this scenario is provided from a renormalization point of view in . The first example of the PC transition was reported by Grassberger et al., who studied probabilistic cellular automata. These 1d models involve the processes $`k3k`$ and $`2k0`$ ($`k`$ stands for kink), very similar to BARW with $`n=2`$. The two-component interacting monomer-dimer model introduced by Kim and Park represents a more complex system with a PC type transition. Other representatives of this class are the three-species monomer-monomer models of , and a generalized Domany-Kinzel SCA , which has two absorbing phases and an active one. A class of general nonequilibrium kinetic Ising models (NEKIM) with combined spin flip dynamics at $`T=0`$ and Kawasaki spin exchange dynamics at $`T=\mathrm{}`$ has been proposed by one of the authors in which, for a range of parameters of the model, a PC-type transition takes place. This model has turned out to be very rich in several respects. As compared to the Grassberger CA models, in NEKIM the rates of random walk, annihilation and kink-production can be controlled independently. It also offers the simplest possibility of studying the effect of a transition which occurs on the level of kinks upon the behaviour of the underlying spin system. The present review is intended to give a summary of the results obtained via computer simulations of the critical properties of NEKIM . Dynamical scaling and a generalized mean field approximation scheme have been applied in the interpretation. Some new variants of NEKIM are also presented here, together with new high-precision data for some of the critical exponents. ## II Non-equilibrium kinetic Ising model (NEKIM) ### A 1d Ising model Before going into the details of NEKIM, let us recall some well known facts about the one-dimensional Ising model. It is defined on a chain of length $`L`$; the Hamiltonian in the presence of a magnetic field $`H`$ is given by $`=3DJ_is_is_{i+1}H_is_i`$, ($`s_i=\pm 1`$). The model does not have a phase transition in the usual sense, but when the temperature $`T`$ approaches zero a critical behaviour is shown. Specifically, $`p_T=e^{\frac{4J}{kT}}`$ plays the role of $`\frac{TT_c}{T_c}`$ in 1d, and in the vicinity of $`T=0`$ critical exponents can be defined as powers of $`p_T`$. Thus, e.g., the critical exponent $`\nu `$ of the coherence length is given through $`\xi p_{T}^{}{}_{}{}^{\nu }`$. Concerning statics, from exact solutions, keeping the leading-order terms for $`T0`$, the critical exponents of ($`k_BT`$ times) the susceptibility, the coherence length and the magnetisation are known to be $`\gamma _s=\nu =1/2`$, and $`\beta _s=0`$, respectively. We can say that the transition is of first order. Fisher’s static scaling law $`\gamma _s=d\nu 2\beta _s`$ is valid. The Ising model possesses no intrinsic dynamics; with the simplest version of Glauber kinetics (see later) it is exactly solvable and the dynamic critical exponent $`Z`$ defined via the relaxation time of the magnetization as $`\tau =\tau _0\xi ^Z`$ has the value $`Z=2`$. (Here $`\tau _0`$ is a characteristic time of order unity.) Turning back to statics, if the magnetic field differs from zero the magnetization is also known exactly; at $`T=0`$ the solution gives: $$m(T=0,H)=sgn(H).$$ (1) Moreover, for $`\xi 1`$ and $`H/kT1`$ the exact solution reduces to $$m2h\xi ;h=H/k_BT.$$ (2) In scaling form one writes: $$m\xi ^{\frac{\beta _s}{\nu }}g(h\xi ^{\frac{\mathrm{\Delta }}{\nu }})$$ (3) where $`\mathrm{\Delta }`$ is the static magnetic critical exponent. Comparison of eqs. (2) and (3) results in $`\beta _s=0`$ and $`\mathrm{\Delta }=\nu `$ . It is clear that the transition is discontinuous at $`H=0`$ as well \[i.e., upon changing the sign of $`H`$ in Eq. (1)\]. In the following the order of limits will always be meant as: (1) $`H0`$, and then (2) $`T0`$. ### B Definition of the model A general form of the Glauber spin-flip transition rate in one-dimension for spin $`s_i`$ sitting at site $`i`$ is ($`s_i=\pm 1`$): $$W_i=\frac{\mathrm{\Gamma }}{2}\left(1+\delta s_{i1}s_{i+1}\right)\left(1\frac{\gamma }{2}s_i\left(s_{i1}+s_{i+1}\right)\right)$$ (4) where $`\gamma =\mathrm{tanh}2J/kT`$ , with $`J`$ denoting the coupling constant in the ferromagnetic Ising Hamiltonian, $`\mathrm{\Gamma }`$ and $`\delta `$ are further parameters which can, in general, also depend on temperature. (Usually the Glauber model is understood as the special case $`\delta =0`$, $`\mathrm{\Gamma }=1`$, while the Metropolis model is obtained by choosing $`\mathrm{\Gamma }=3/2`$, $`\delta =1/3`$). There are three independent rates: $`w_{same}={\displaystyle \frac{\mathrm{\Gamma }}{2}}\left(1+\delta \right)\left(1\gamma \right)`$ (5) $`w_{oppo}={\displaystyle \frac{\mathrm{\Gamma }}{2}}\left(1+\delta \right)\left(1+\gamma \right)`$ (6) $`w_{indif}={\displaystyle \frac{\mathrm{\Gamma }}{2}}\left(1\delta \right),`$ (7) where the subscripts $`same`$, etc., indicate the three possible neighborhoods of a given spin($`,`$ and $``$,respectively). In the following $`T=0`$ will be taken, thus $`\gamma =1`$, $`w_{same}=0`$ and $`\mathrm{\Gamma }`$, $`\delta `$ are parameters to be varied. The Kawasaki spin-exchange transition rate of neighbouring spins is: $$w_{i,i+1}(s_i,s_{i+1})=\frac{p_{ex}}{2}\left(1s_is_{i+1}\right)\left[1\frac{\gamma }{2}\left(s_{i1}s_i+s_{i+1}s_{i+2}\right)\right].$$ (8) At $`T=\mathrm{}`$ ($`\gamma =0`$) the above exchange is simply an unconditional nearest neighbor exchange: $$w_{i,i+1}=\frac{1}{2}p_{ex}\left[1s_is_{i+1}\right]$$ (9) where $`p_{ex}`$ is the probability of spin exchange. The transition probabilities in Eqs. (4) and (9) are responsible for the basic elementary processes of kinks. Kinks separating two ferromagnetically ordered domains can carry out random walks with probability $$p_{rw}2w_{indif}=\mathrm{\Gamma }\left(1\delta \right)$$ (10) while two kinks at neighbouring positions will annihilate with probability $$p_{an}w_{oppo}=\mathrm{\Gamma }\left(1+\delta \right)$$ (11) ($`w_{same}`$ is responsible for creation of kink pairs inside of ordered domains at $`T0`$). In case of the spin exchanges, which also act only at domain boundaries, the process of main importance here is that a kink can produce two offspring at the next time step with probability $$p_{k3k}p_{ex}.$$ (12) The above-mentioned three processes compete, and it depends on the values of the parameters $`\mathrm{\Gamma }`$, $`\delta `$ and $`p_{ex}`$ what the result of this competition will be. It is important to realize that the process $`k3k`$ can develop into propagating offspring production only if $`p_{rw}>p_{an}`$, i.e., the new kinks are able to travel on the average some lattice points away from their place of birth, and can thus avoid immediate annihilation. It is seen from the above definitions that $`\delta <0`$ is necessary for this to happen. In the opposite case the only effect of the $`k3k`$ process on the usual Ising kinetics is to soften domain walls. This heuristic argument is supported by simulations as well as theoretical considerations, as will be discussed in the next section. ### C Phase boundary of the PC transition in NEKIM In all of our investigations we have considered a simplified version of the above model, keeping only two parameters instead of three by imposing the normalization condition $`p_{rw}+p_{an}+p_{k3k}=1`$, i.e., $$2\mathrm{\Gamma }+p_{ex}=1.$$ (13) In the plane of parameters $`p_{ex}`$ and $`\delta =\frac{p_{rw}/p_{an}1}{p_{rw}/p_{an}+1}`$ the phase diagram, obtained by computer simulation, is as shown in Fig. 1. There is a line of second-order phase transitions, the order parameter being the density of kinks. The phase boundary was obtained by measuring $`\rho \left(t\right)`$, the density of kinks, starting from a random initial distribution, and locating the phase transition points via $`\rho \left(t\right)t^\alpha `$ with $`\alpha =.28\pm .01`$. The lattice was typically $`L=2000`$ sites; we generally averaged over 2000 independent runs. The dotted vertical line in Fig. 1 indicates the critical point that we investigated in greater detail; further critical characteristics were determined along this line. The line of phase transitions separates two kinds of steady states reachable by the system for large times: in the Ising phase, supposing that an even number of kinks are present in the initial states, the system orders in one of the possible ferromagnetic states of all spins up or all spins down, while the active phase is disordered, from the point of view of the underlying spins, due to the steadily growing number of kinks. For determining the phase boundary, as well as other exponents discussed below, we used the scaling considerations of Grassberger . The density $`\rho (x,t,ϵ)`$ of kinks was assumed to follow the scaling form $$\rho (x,t,ϵ)t^\alpha \varphi (ϵx^{1/\nu _{}},ϵt^{1/\nu _{}})$$ (14) Here $`ϵ`$ measures the deviation from the critical probability at which the branching transition occurs, $`\nu _{}`$ and $`\nu _{}`$ are exponents of coherence lengths in space and time directions, respectively. By definition, $`\nu _{}=\nu _{}Z`$, where $`Z`$ is the dynamic critical exponent. $`\varphi (a,b)`$ is analytic near $`a=0`$ and $`b=0`$. Using Eq. (14) the following relations can be deduced. When starting from a random initial state the exponent $`\beta `$ characterizes the growth of the average kink density in the active phase: $$\rho \left(ϵ\right)=\underset{t\mathrm{}}{lim}\rho (x,t,ϵ)ϵ^\beta $$ (15) while the decrease of density at the critical point is given by $$\rho \left(t\right)t^\alpha .$$ (16) The exponents are connected by the scaling law: $$\beta =\nu _{}\alpha =\nu _{}Z\alpha $$ (17) The phase boundary shown in Fig 1 was identified using Eq. (16) . The initial state was random, with zero average magnetization. At the critical point marked on the phase boundary we recently obtained the value $`\alpha =.280\left(5\right)`$. This point was chosen in a region of the parameters $`(p_{ex},\delta )`$, where the effect of transients in time-dependent simulations is small. Critical exponents were measured around the point: $`\mathrm{\Gamma }=0.35`$, $`\delta _c=0.395\pm .001`$ and $`p_{ex}=0.3`$. The deviation from the critical point $`ϵ`$ was taken in the $`\delta `$-direction. It is worth noting that in all of our simulations of the NEKIM rule, spin-flips and spin-exchanges alternated at each time step. Spin-flips were implemented using two-sublattice updating, while L exchange attempts were counted as one time-step for exchange updating. The numerical value of the phase transition point is sensitive to the manner of updating (thus, e.g., with sequential updating in both processes, $`\delta _c`$ decreases by more than ten percent at the same values of $`\mathrm{\Gamma }`$ and $`p_{ex}`$). Concerning $`\beta `$, we carried out quite recently a high-precision numerical measurement; the result is shown in Fig. 2. To see the corrections to scaling as well, we determined the effective exponent $`\beta _{eff}\left(ϵ\right)`$, which is defined as the local slope of $`\rho \left(ϵ\right)`$ in a log-log representation between the data points $`(i1,i)`$ $$\beta _{eff}\left(ϵ_i\right)=\frac{\mathrm{ln}\rho \left(ϵ_i\right)\mathrm{ln}\rho \left(ϵ_{i1}\right)}{\mathrm{ln}ϵ_i\mathrm{ln}ϵ_{i1}},$$ (18) where $`ϵ_i=\delta _c\delta _i`$, providing an estimate $$\beta =\underset{ϵ0}{lim}\beta _{eff}\left(ϵ\right).$$ (19) As shown in Fig. 2, the effective exponent increases slowly and tends towards the expected PC value as $`ϵ0`$. A simple linear extrapolation yields the estimate $`\beta =0.95\left(2\right)`$. The transients mentioned above in connection with the time-dependent simulations show up for small times and are particularly discernible at the two ends of the phase boundary, where the spin-flip and spin-exchange processes have very different time scales. For this reason, near $`p_{ex}=1.0`$, where we get close to $`\delta =0`$ (the Glauber case), it is very difficult to get reliable numerical estimates, even for $`\alpha `$; the exponent has been found to grow slowly with time, from a value near zero at small times. It is important to notice that according to Eq. (13), $`p_{ex}=1`$ is approached together with $`\mathrm{\Gamma }0`$; thus $`\frac{p_{ex}}{\mathrm{\Gamma }}\mathrm{}`$, as otherwise $`\delta _c=0`$ cannot be reached. Several runs with different parameter values \[including cases without the restriction, Eq. (13)\], have been performed in this Glauber limit; all show that the $`\delta =0`$ case remains Ising-like for all values of $`\mathrm{\Gamma }`$ and $`p_{ex}`$: the exponent $`\alpha `$ tends to $`0.5`$ for late times. We therefore conjectured that the asymptote of the phase boundary is $`\delta =0`$ for $`p_{ex}=1.0`$, and thus that $`p_{rw}>p_{an}`$ is a necessary condition for a PC-type phase transition. In a recent paper, using exact methods, Mussawisade, Santos and Schütz confirmed the above conjecture. These authors study a one-dimensional BARW model with an even number of nearest-neighbour offspring. They start with spin kinetics; in their notation D, $`\lambda `$ and $`\alpha `$ are the rates of spin diffusion, annihilation and spin-exchange, respectively, and transform to a corresponding particle (kink) kinetics to arrive at a BARW model. (In their notation $`D=p_{RW}`$, $`\lambda =p_{an}`$ and $`\alpha =2p_{ex}`$.) Using standard field-theoretic techniques, Mussawisade et al. derived duality relations $`\stackrel{~}{\lambda }=\lambda ,`$ (20) $`\stackrel{~}{D}=\lambda +\alpha ,`$ (21) $`\stackrel{~}{\alpha }=D\lambda `$ (22) which map the phase diagram onto itself in a nontrivial way and divide the parameter space into two distinct regions separated by the self-dual line : $$D=\lambda +\alpha .$$ (23) The regions mapped onto each other have, of course, the same physical properties. iN particular, the line $`\alpha =0`$ maps onto the line $`D=\lambda `$ and the fast-diffusion limit to the limit $`\alpha \mathrm{}`$. In the fast diffusion limit the authors find that the system undergoes a mean-field transition, but with particle-number fluctuations that deviate from those given by the MF approximation. So the nature of the phase transition in this limit is not PC but MF. The observation that for large $`D`$, any small $`\alpha `$ brings the system into the active phase, together with duality, predicts a phase transition at $`D=\lambda `$, in the limit $`\alpha \mathrm{}`$. The exact result of confirms the conjecture - stemming from our simulations - on the location of the phase transition point in the phase diagram of NEKIM in the limit $`p_{ex}1`$, which is indeed at $`\delta =0`$. The dual limit $`D\mathrm{}`$ covers the neighbourhood of the point $`\delta =1`$, $`p_{ex}=0`$. This result predicts an infinite slope of the phase boundary in this representation, at this point, as is also apparent in Fig.1 It is important to notice that the duality transformation does not preserve the normalization condition of NEKIM, eq(13), used in all of our numerical studies. The self-dual line of $`D=\lambda +\alpha `$,which in terms of our parametrization reads $$\delta =\frac{2p_{ex}}{1p_{ex}}$$ (24) does not conserve the above-mentioned normalization condition either. In the phase diagram of Fig. 1 it corresponds to a line which starts at $`(0,0)`$ and ends at $`(1/3,1)`$ in the $`(p_{ex},\delta )`$ plane. Finally, in ref. a proportionality relation between the (time-dependent) kink density for a half-filled random initial state, and the survival probability of two single particles in an initially otherwise empty system has also been derived. Concerning the respective critical exponents $`\alpha `$ and $`\delta `$ (the latter being defined through the critical behaviour of the survival probability $`P\left(t\right)t^\delta `$), they have been shown to coincide within the error of numerical simulations in , as required by scaling. ### D Long-range initial conditions The symmetry between the two extreme inital cases in time dependent simulations (single seed versus homogeneous random state) concerning the kink density decay inspired us to investigate initial conditions with long-range two-point correlations in the NEKIM model. This procedure was shown to cause continuously changing density decay exponents in case of DP transitions . In this way it has been possible to interpolate continuously between the two extreme initial cases. In the present case we have started from initial states containing an even number of kinks to ensure kink number conservation mod 2 and possessing kink-kink correlations of the form $`<k_ik_{i+x}>x^{\left(1\sigma \right)}`$ with $`\sigma `$ lying in the interval $`(0,1)`$. Here $`k_i`$ denotes a kink at site $`i`$. These states have been generated by the same serial algorithm as described and numerically tested in ref. . The kink density in surviving samples was measured in systems of $`L=12000`$ sites, for up to $`t_{max}=80000`$ time steps; we observe good quality of scaling over a time interval of three decades $`(80,80000)`$ (see Fig. 3) . As one can see, the kink density increases with exponent $`\alpha 0.28`$ for $`\sigma =0`$, where in principle only one pair of kinks is placed on the lattice owing to the very short range correlations. Without the restriction to surviving samples we would have expected a constant, the above exponent corresponds to dividing the kink density by the survival probability,$`P\left(t\right)t^\delta `$ and the above value is in agreement with our former simulation results for $`\delta `$ . In the other extreme case for $`\sigma =1`$ the kink density decays with with an exponent $`\alpha 0.28`$, again in agreement with our expectations in case of a random initial state . In between the exponent $`\alpha `$ changes linearly as a function of $`\sigma `$ and changes its sign at $`\sigma =0.5`$. This means that the state generated with $`\sigma =1/2`$ is very near to the situation which can be reached for $`t\mathrm{}`$, i.e., the steady-state limit. We mentioned at the end of the last section that the duality transformation of ref. predicts the equality of the above two extremal exponents (though it also follows from scaling). One can make the conjecture that in greater generality the duality transformation may connect also cases with initial two-point correlation exponents: $`\sigma (1\sigma )`$. ### E Variants of the NEKIM model #### 1 Varying exchange-range model Now we generalize the original NEKIM model by allowing the range of the spin-exchange to vary. Namely, Eq. (9) is replaced by $$w_{i,i+k}=\frac{1}{2}p_{ex}\left[1s_is_{i+k}\right],$$ (25) where $`i`$ is a randomly chosen site and $`s_i`$ is allowed to exchange with $`s_{i+k}`$ with probability $`p_{ex}`$. Site $`k`$ is again randomly chosen in the interval $`1R`$ , $`R`$ being thus the range of exchange. The spin-flip part of the model will be unchanged. We have carried out numerical studies with this generalized model in order to locate the lines of Ising-to-active PC-type phase transitions. The method of updating was as before. It is worth mentioning, that besides $`k3k`$, the process $`k5k`$ can also occur for $`R3`$, and the new kink pairs are not necessarily neighbors. The character of the phase transition line at $`R>1`$ is similar to that for $`R=1`$, except that the active phase extends, asymptotically, down to $`\delta _c=0`$. This is illustrated in Fig. 4, where besides $`R=1`$, the case $`R=3`$ is also depicted: the critical value of $`\delta _c`$ is shown as a function of $`p_{ex}`$ \[curves a) and b)\]. Moreover, $`\delta _c`$ as a function of $`R`$ is also shown at constant $`\mathrm{\Gamma }=.35`$ . The abscissa has been suitably chosen to squeeze the whole (infinite) range of $`R`$ between $`0`$ and $`1`$ and for getting phase lines of comparable size (hence the power 4 of $`R/\left(1+R\right)`$ in case of curve c)). The phase boundaries were obtained by measuring $`\rho \left(t\right)`$, the density of kinks, starting from a random initial distribution and locating the phase transition points by $`\rho \left(t\right)t^\alpha `$ with $`\alpha =0.27\pm .04`$. We typically used a system of $`L=2000`$ sites and averaged over 500 independent runs. Besides the critical exponents we have also determined the change of $`\delta _c`$ with $`1/R`$ numerically at fixed $`\mathrm{\Gamma }=.35,p_{ex}=.3`$ . Over the decade of $`R=440`$ we have found $$\delta _c2.0\left(1/R\right)^2,$$ (26) which is reminiscent of a crossover type behaviour of equilibrium and non-equilibrium phase transitions , here with crossover exponent $`1/2`$. It should be noted here, that to get closer to the expected $`\delta _{cMF}=0`$, longer chains (we used $`l`$ values up to $`20000`$) and longer runs (here up to $`t=510^4`$) would have been necessary. The former to ensure $`lR`$ and the latter to overcome the long transients present in the first few decades of time steps. In what follows we will always refer to the MF limit in connection with $`p_{ex}1`$ (i.e., $`p_{ex}/\mathrm{\Gamma }\mathrm{}`$), for the sake of concreteness, but keep in mind that $`R\mathrm{}`$ can play the same role. #### 2 Cellular automaton version of NEKIM Another generalization is the probabilistic cellular automaton version of NEKIM, which consists in keeping the rates given in eqs.(7) and prescribing synchronous updating. Vichniac investigated the question how well cellular automata simulate the Ising model with the conclusion that the unwanted feedback effects present at finite temperatures are absent at $`T=0`$ and the growth of domains is enhanced; equilibrium is reached quickly. In the language of Glauber kinetics, this refers to the $`\delta 0`$ case. With synchronous updating, it turns out that the $`T=0`$ Glauber spin-flip part of NEKIM itself leads to processes of the type $`k3k`$, and for certain values of parameter-pairs $`(\mathrm{\Gamma },\delta )`$ with $`\delta <0`$, a PC-type transition takes place. For $`\mathrm{\Gamma }=.5`$ the critical value of $`\delta `$ is $`\delta _c=.425\left(5\right)`$ , while e.g., for $`\mathrm{\Gamma }=.35,\delta _c=.535\left(5\right)`$. For $`\delta >\delta _c`$ the phase is the active one,while in the opposite case it is absorbing. All the characteristic exponents can be checked with a much higher speed, than for the original NEKIM. The phase boundary of the NEKIM-CA in the $`(\mathrm{\Gamma },\delta )`$ plane is similar to that in Fig. 1 except for the highest value of $`\mathrm{\Gamma }=1`$ $`\delta _c=0`$ cannot be reached; the limiting value is $`\delta _c=.065`$. For the limit $`\delta =0`$ Vichniac’s considerations apply, i.e., it is an Ising phase. The limiting case $`\mathrm{\Gamma }0`$ with $`\delta _c1`$ is very hard from the point of view of simulations, moreover the case $`\mathrm{\Gamma }=0`$ is pathological: the initial spin-distribution freezes in. The corresponding phase diagram is shown in Fig. 5. #### 3 NEKIM in a magnetic field It has been shown earlier, in the framework of a model different from NEKIM exhibiting a PC-type phase transition , that upon breaking the $`Z_2`$ symmetry of the absorbing phases, DP universality is recovered. In case of the Glauber model a similar situation arises with the introduction of a magnetic field. In the presence of a magnetic field the Glauber transition rates given in Sec. 2 are modified so: $`w_{indif}^h`$ $`=`$ $`w_{indif}\left(1hs_i\right),`$ (27) $`w_{oppo}^h`$ $`=`$ $`w_{oppo}\left(1hs_i\right),`$ (28) $`h`$ $`=`$ $`\mathrm{tanh}\left({\displaystyle \frac{H}{k_BT}}\right).`$ (29) Fig. 6 shows the phase diagram of NEKIM in the $`(h,\delta )`$ plane , starting at the reference PC point for $`h=0`$ used in this paper ($`\delta _c=.395`$). We have applied only random initial state simulations to find the points of the line of phase transitions (critical exponents: $`\alpha =.17\left(2\right)`$, $`\beta =.26\left(2\right)`$, corresponding to the DP universality class, as expected.) It is seen, that with increasing field strength the critical point is shifted to more and more negative values of $`\delta `$. Finally, it is worth mentioning a further possible varint of NEKIM \- again without a magnetic field - in which the Kawasaki rate Eq. (8) is considered at some finite temperature, instead of $`T=\mathrm{}`$, but keeping $`T=0`$ in the Glauber dynamics. As lowering the temperature of the spin-exchange process acts against kink production, the result is that the active phase shrinks. For further details see Ref. . ## III Generalized mean-field theory and Coherent anomaly extrapolation Using the CA version of NEKIM described in the previous section we performed $`n`$-site cluster mean-field calculations (GMF) up to $`n=6`$, and a coherent anomaly extrapolation (CAM), resulting in numerical estimates for critical exponents in fairly good agreement with those of other methods. With this method we could also handle a symmetry-breaking field that favors the spin-flips in one direction. For this SCA model we applied the generalized mean-field technique introduced by Gutowitz and Dickman for nonequilibrium statistical systems. This is based on the calculation of transitions of $`n`$-site cluster probabilities $`P_n\left(\left\{s_i\right\}\right)`$ as $$P_n^{t+1}(s_1,\mathrm{},s_n)=\left(\left\{P_m^t(s_1,\mathrm{},s_m)\right\}\right)=P_n^t(s_1,\mathrm{}s_n)$$ (30) where $``$ is a function depending on the update rules and the last equation refers to our solution in the steady state limit. At the $`k`$-point level of approximation, correlations are neglected for $`n>k`$, that is, $`P_n(s_1,\mathrm{},s_n)`$ is expressed by using the Bayesian extension process . $$P_n(s_1,\mathrm{},s_n)=\frac{\underset{j=0}{\overset{j=nk}{}}P_k(s_{1+j},\mathrm{},s_{k+j})}{_{j=1}^{j=nk}P_{k1}(s_{1+j},\mathrm{},s_{k1+j})}.$$ (31) The number of independent variables grows more slowly than $`2^n1`$, owing to internal symmetries (reflection, translation) of the update rule and the ‘block probability consistency’ relations: $`P_n(s_1,\mathrm{},s_n)`$ $`=`$ $`{\displaystyle \underset{s_{n+1}}{}}P_{n+1}(s_1,\mathrm{},s_n,s_{n+1}),`$ (32) $`P_n(s_1,\mathrm{},s_n)`$ $`=`$ $`{\displaystyle \underset{s_0}{}}P_{n+1}(s_0,s_1,\mathrm{},s_n).`$ (33) The series of GMF approximations can now provide a basis for an extrapolation technique. One such a method is the coherent anomaly method (CAM) introduced by Suzuki . According to the CAM (based on scaling) the GMF solution for a physical variable $`A`$ (example the kink density) at a given level ($`n`$) of approximation – in the vicinity of the critical point, $`\delta _c`$ – can be expressed as the product of the mean-field scaling law multiplied by the anomaly factor $`\overline{a}_n`$: $$A_n=\overline{a}_n\left(\delta /\delta _c^n1\right)^{\varphi _{MF}},$$ (34) and the $`n`$-th approximation of critical exponents of the true singular behavior ($`\varphi _n`$) can be obtained via the scaling behavior of anomaly factors: $$\overline{a}_n\mathrm{\Delta }_n^{\varphi _n\varphi _{MF}}$$ (35) where we have used the $`\mathrm{\Delta }_n=\left(\delta _c/\delta _c^n\delta _c^n/\delta _c\right)`$ invariant variable instead of $`ϵ`$, that was introduced to make the CAM results independent of using $`\delta _c`$ or $`1/\delta _c`$ coupling (). Here $`\delta _c`$ is the critical point estimated by extrapolation from the sequence approximations $`\delta _c^n`$. Since we can solve the GMF equations for $`n6`$, we have taken into account corrections to scaling, and determined the true exponents with the nonlinear fitting form : $$\overline{a}_n=b\mathrm{\Delta }_n^{\varphi _n\varphi _{MF}}+c\mathrm{\Delta }_n^{\varphi _n\varphi _{MF}+1}$$ (36) where $`b`$ and $`c`$ are coefficients to be varied. As discussed in a previous section, by duality symmetry the $`p_{ex}=0`$ axis is mapped into the $`\delta =0`$ axis and the neighborhood of the ($`p_{ex}=1,\delta =0`$) tricritical point which can be well described by the mean-field approximations to the ($`p_{ex}=0,\delta =1`$) point. Therefore this point is also a tricritical point and the mean-field results are applicable in the neighborhood. ### A $`Z_2`$ symmetric case First we show the results of the GMF+CAM calculations without external field. For $`n=1`$, we have the traditional mean-field equation for the magnetization: $$dm/dt=\delta \mathrm{\Gamma }m\left(m^21\right)$$ (37) and for the kink density : $$d\rho /dt=2\delta \mathrm{\Gamma }\rho \left(13\rho +2\rho ^2\right)$$ (38) The stable, analytic solution of the kink density exhibits a jump at $`\delta =0`$. $`\rho _1(\mathrm{})=\{\begin{array}{cc}\hfill 1/2:& \delta <0\hfill \\ \hfill 0:& \delta 0\hfill \end{array}`$ The $`n=2`$ approximation gives for the density of kinks, $`\rho \left(\mathrm{}\right)`$, the following expression: $$\rho _2\left(\mathrm{}\right)=\frac{\frac{3}{4}p_{rw}^{}{}_{}{}^{2}+p_{an}p_{rw}p_{an}\sqrt{\frac{1}{16}p_{rw}^{}{}_{}{}^{4}+\frac{3}{2}p_{rw}^{}{}_{}{}^{2}p_{an}\frac{1}{2}p_{rw}^{}{}_{}{}^{3}p_{an}+p_{an}^{}{}_{}{}^{2}2p_{rw}p_{an}^{}{}_{}{}^{2}}}{2\left(\frac{1}{2}p_{rw}^{}{}_{}{}^{2}p_{rw}p_{an}+p_{an}^{}{}_{}{}^{2}\right)}$$ (39) for $`\delta <0`$. For $`\delta >0`$ $`\rho \left(\mathrm{}\right)=0`$, i.e., GMF still predicts a first-order transition for $`\delta =0`$; the jump size in $`\rho \left(\mathrm{}\right)`$ at $`\delta =0`$, however, decreases monotonically with decreasing $`\mathrm{\Gamma }`$, according to eq. (39). The GMF equations can only be solved numerically for $`n>2`$. We determined the solutions of the $`n=3,4,5,6`$ approximations for the kink density at i). $`\mathrm{\Gamma }=0.35`$ (Fig.7 a) and of the $`n=3,4,5`$ approximations at ii). $`\mathrm{\Gamma }=0.05`$ (Fig. (7 b)). As we can see, the transition curves become continuous, with negative values for $`\delta _c^n`$ ($`\delta _c^n`$ denotes the value of $`\delta `$ in the $`n`$-th approximation for which the corresponding $`\rho \left(\mathrm{}\right)`$ becomes zero). Moreover, $`\delta _c^n`$ increases with growing $`n`$ values. As increasing $`n`$ corresponds to decreasing mixing, i.e., decreasing $`p_{ex}`$, the tendency shown by the above results is correct. Figs. 8 a) and 8 b) show a quantitative - though only tentative - comparison between the results of GMF and the simulated NEKIM phase diagrams. The obtained GMF data for $`\delta _c^n`$, corresponding to $`n=3,4,\mathrm{},6`$ ($`\mathrm{\Gamma }=0.35`$) are depicted in Fig. 8 a) as a function of $`1/\left(n3\right)`$, together with results of simulations. The correspondence between $`n`$ and $`p_{ex}`$ was chosen as the simplest conceivable one. (Note that $`\delta _c0`$ is obtained first for $`n=4`$.) The simulated phase diagram was obtained without requiring the normalization condition, Eq. (13), at fixed $`\mathrm{\Gamma }=0.35`$. In this case, the $`\delta _c=0`$ limit, of course, is not reached and a purely second-order phase transition line can be compared with the GMF results (for $`n`$ values where the latter also predicts a second-order transition). Simulations for $`R=3`$ were found to lead to $`\delta _c`$ values low enough to fit GMF data. The (polynomial) extrapolation of GMF data to $`n\mathrm{}`$ (corresponding to $`p_{ex}=0`$ , i.e., plain spin-flip), also shown in Fig. 8, could be expected to approach $`\delta =1`$. That this is not case may be due to the circumstances that upon increasing $`n`$ i) GMF starts here from a first-order MF phase transition, which ii) becomes second-order, and iii) GMF should end up at the dual point discussed in Sec. II.C, while this symmetry is not included in the applied approximation procedure. Fig. 8 (b) shows the $`n=3,4,5`$ results for $`\mathrm{\Gamma }=0.05`$, which are compared now with the $`R=1`$ simulation data. The results of our GMF approximation useful for a CAM analysis are the $`n=4,5,6`$ data, while the $`n=3`$ result can be taken to represent the lowest-order MF approximation (with $`\delta _{c}^{}{}_{}{}^{MF}=0`$) for a continuous transition (no jump in $`\rho _3`$). For $`\delta _c`$ we use a polynomial extrapolation. In the $`n=3`$ approximation the exponent $`\beta =1.0064`$, thus $`\beta _{MF}1`$. Consequently, as the table below shows, the anomaly factor does not depend on $`n`$. This means, according to Eq. (35), that the exponent is estimated to be equal to the MF value $`\beta \beta _{MF}=1`$. ### B Broken $`Z_2`$ symmetry case The transition probabilities of NEKIM, in the presence of en external magnetic field H have been given in eqs.(27)-(29), in the previous section. Now we extend the method for the determination of the exponent of the order-parameter fluctuation as well : $$\chi \left(ϵ\right)=L\left(<\rho ^2><\rho >^2\right)ϵ^{\gamma _n}.$$ (40) The traditional mean-field solution ($`n=1`$) results in stable solutions for the magnetization : $`m_1=`$ $`{\displaystyle \frac{h}{\delta }},`$ $`if\delta <0\mathrm{and}h^2/\delta ^2<1`$ (41) $`m_1=`$ $`\mathrm{sgn}\left(h\right),`$ $`\mathrm{otherwise}.`$ (42) and for the kink-concentration : $`\rho _1\left(\mathrm{}\right)=`$ $`{\displaystyle \frac{1}{2}}\left(1\left({\displaystyle \frac{h}{\delta }}\right)^2\right),`$ $`if\delta <0\mathrm{and}h^2/\delta ^2<1`$ (43) $`\rho _1\left(\mathrm{}\right)=`$ $`0,`$ $`\mathrm{otherwise}.`$ (44) For $`n>1`$ the solutions can again only be found numerically. Increasing the order of approximation, the critical point estimates $`\delta _c^n`$ shift to more negative values similarly to the $`H=0`$ case. The $`lim_n\mathrm{}\delta _c^n\left(h\right)`$ values have been determined with quadratic extrapolation for $`h=0.01,0.05,0.08`$, and 0.1. The resulting curves for $`\rho _n\left(\delta \right)`$ and $`\chi _n\left(\delta \right)`$ are shown in Figs. 9 (a) and (b), respectively, for the case of $`h=0.1`$ in different orders $`n`$ of the GMF approximation. Naturally, these curves exhibit a mean-field type singularity at the critical point : $`\rho _n`$ $`\overline{\rho }_n\left(\delta /\delta _c^n1\right)^{\beta _{MF}}`$ (45) $`\chi _n`$ $`\overline{\chi }_n\left(\delta /\delta _c^n1\right)^{\gamma _{MF}},`$ (46) with $`\beta _{MF}=1`$ and $`\gamma _{MF}=1`$. The results of the CAM extrapolation for various $`h`$-values are shown in the table below : As to the case $`h=0`$, we could not determine the exponent $`\gamma _n`$ because the low level GMF calculations resulted in discontinuous phase transition solutions – which we cannot use in the CAM extrapolation – and so we had too few data points to achieve a stable non-linear fitting. Higher-order GMF solutions would help, but that requires the solution of a set of nonlinear equations with more than $`72`$ independent variables. This problem does not occur for $`h0`$; the above results - being based on the full set of approximations ($`n=1,\mathrm{},6`$) - are fairly stable. ## IV Domain growth versus cluster growth properties; Hyperscaling In the field of domain-growth kinetics it has long been accepted that the scaling exponent of $`L\left(t\right)`$, the characteristic domain size, is equal to $`1/2`$ if the order parameter is not conserved. In the case of a 1d Ising spin chain of length $`L`$, the structure factor at the ferromagnetic Bragg peak is defined as: $`S(0,t)=L\left[<m^2><m>^2\right],m=\frac{1}{L}_is_i`$, ($`s_i=\pm 1`$). If the conditions of validity of scaling are fulfilled then $$S(0,t))[L\left(t\right)]^d;L(t)t^x$$ (47) where now $`d=1`$ and $`x=1/2`$. Another quantity usually considered is the excess energy $`\mathrm{\Delta }E\left(t\right)=E\left(t\right)E_T`$ ( $`E_T`$ is the internal energy of a monodomain sample at the temperature of quench), which in our case is proportional to the kink density $`\rho \left(t\right)=\frac{1}{L}<_i\frac{1}{2}\left(1s_is_{i+1}\right)>`$. $$\rho \left(t\right)\frac{1}{L\left(t\right)}t^y$$ (48) with $`y=1/2`$ in the Glauber-Ising case, expressing the well known dependence on time of annihilating random walks ($`y=\alpha `$ in the notation of section 2.). We also measured $`S\left(t\right)`$ at the PC point and found power-law behaviour with an exponent $`x=.575\left(5\right)`$ . The second characteristic growth length at the PC point is the cluster size, defined through the square-root of $`<R^2\left(t\right)>t^z`$. The latter is obtained by starting either from two neighbouring kink initial states (see e.g.), or from a single kink . Both length-exponents are, however connected with $`Z`$, the dynamic critical exponent, since at the PC transition the only dominant length is the (time-dependent) correlation length. It was shown in and that $`\frac{z}{2}=\frac{1}{Z}`$ follows from scaling for one-kink and two-kink initial states, respectively. In our simulations we also found that $`x=1/Z`$, within numerical error, at the PC point, and thus $$x=z/2$$ (49) follows. Exponents of cluster growth are connected by a hyperscaling relation first established by Grassberger and de la Torre for the directed percolation transition. It does not apply to the PC transition in the same form , where the dependence on the initial state (one or two kinks) manifests itself in two cluster-growth quantities: the kink-number $`N(t,ϵ)t^\eta f\left(ϵt^{\frac{1}{\nu _{}Z}}\right)`$ and the survival probability $`P(t,ϵ)t^\delta g\left(ϵt^{\frac{1}{\nu _{}Z}}\right)`$ (This $`\delta `$ has, of course, nothing to do with the parameter $`\delta `$ of NEKIM). In systems with infinitely many absorbing states, where $`\delta `$ and $`\eta `$ depend on the density of particles in the initial configuration, a generalised scaling relation was found to replace the original relation, $`2\delta +\eta =z/2`$, namely : $$2\left(\beta ^{^{}}+\beta \right)\frac{1}{\nu _{}Z}+2\eta ^{^{}}=z^{^{}}$$ (50) which is also applicable to the present situation. In eq(50) $`\beta ^{^{}}`$ is defined through $`lim_t\mathrm{}P(t,ϵ)ϵ^\beta ^{^{}}`$ and the scaling relation $`\beta ^{^{}}=\delta ^{^{}}\nu _{}Z`$ holds. A prime on an exponent indicates that it may depend on the initial state. For the PC transition the exponent $`z`$ has proven (within error) to be independent of the initial configuration, $`z^{^{}}=z`$. In case of a single kink initial state all samples survive thus $`\delta ^{^{}}=0`$, and via the above scaling relation also $`\beta ^{^{}}=0`$. Eq(50) then becomes: $`\frac{2\beta }{\nu _{}Z}+2\eta ^{^{}}=z`$. For $`t\mathrm{}`$, however, $`lim_t\mathrm{}N(t,ϵ)ϵ^\beta `$ has to hold ( the steady state reached can not depend on the initial state provided samples survive). Thus using the above scaling form for $`N(t,ϵ)`$, $`\eta =\frac{\beta }{\nu _{}Z}`$ follows, and the hyperscaling law can be cast into the form: $$\frac{4\beta }{\nu _{}Z}=z.$$ (51) Starting with a two-kink initial state, however, $`\beta ^{^{}}=\beta `$ and $`\eta ^{^{}}=0`$ holds . With these values Eq. (50) leads again to Eq. (51). As $`z=2/Z`$, Eq. (51) involves only such quantities, which are defined also when starting with a random initial state. Using Eq. (17), Eq. (51) becomes $$2y=1/Z=x$$ (52) In this way the factor of two between $`x`$ and $`y`$ appearing at the PC point between the exponents of a spin-bound and a kink-bound quantity (and which has been found also for a variety of critical exponents, see ) could be explained as following from (hyper)scaling. ## V Effect of the PC transition on the properties of the underlying spin system Time-dependent simulations, finite-size scaling and the dynamic early-time MC method have been applied to investigate the behaviour of the 1d spin system under the effect of the PC transition. We found that within error of the numerical studies, the dynamic critical exponent of the kinks $`Z`$ and that of the spins $`Z_c`$ agree, and have the value: $`Z=Z_c=1.75`$. Thus in comparison with the Glauber-Ising value $`Z=2.0`$, $`Z`$ decreases, as do $`\gamma `$ and $`\nu `$: $`\gamma =\nu =.444`$ (instead of 1/2). The PC point is the endpoint of a line of first-order phase transitions (by keeping $`p_{ex}`$ and $`\mathrm{\Gamma }`$ fixed and changing $`\delta `$ through negative values to $`\delta _c`$), where $`\beta =0`$ still holds, as does Fisher’s scaling law: $`\gamma =d\nu 2\beta `$. To obtain these values the PC point has been approached from the direction of finite temperatures of the spin-flip process, by varying $`p_T=e^{4J/kT}`$. The dynamical persistency exponent $`\mathrm{\Theta }`$ and the exponent $`\lambda `$ characterising the two-time autocorrelation function of the total magnetization under nonequilibrium conditions have also been investigated numerically at the PC point. It has been found that the PC transition has a strong effect: the process becomes non-Markovian and the above exponents exhibit drastic changes as compared to the Glauber-Ising case . These results together with critical exponents obtained earlier in are summarized in Table III. We have also investigated the effect of critical fluctuations at the PC point on the spreading of spins, and found the analogue of compact directed percolation (CDP) to exist. Compact directed percolation is known to appear at the endpoint of the directed percolation critical line of the Domany-Kinzel cellular automaton in $`1+1`$ dimensions. Equivalently, such a transition occurs at zero temperature in a magnetic field $`h`$, upon changing the sign of $`h`$, in the one-dimensional Glauber-Ising model, with well known exponents characterising spin-cluster growth . For the spreading process of a single spin $`s\left(i\right)=1`$ in the sea of downward spins($`s\left(j\right)=1`$ for $`ji`$), the exponents $`\delta _s`$, $`\eta _s`$ and $`z_s`$ are defined at the transition point by the power-laws governing the density of $`1`$’s $`n_st^{\eta _s}`$, the survival probability $`P_s\left(t\right)t^{\delta _s}`$, and the mean-square distance of spreading $`<R_s^2\left(t\right)>t^{z_s}`$. We obtained $`\eta =0`$, $`\delta =1/2`$ and $`z=1`$, in accord with the hyperscaling relation in the form appropriate for compact clusters: $$\eta _s+\delta _s=z_s/2$$ (53) At the PC point we found that the characteristic exponents differ from those of the CDP transition, as could be expected. But basic similarities remain. Thus the transition which takes place upon changing the sign of the magnetic field is of first-order and its exponents satisfy Eq. (53). Accordingly it can be termed as ’compact’, and we call it the compact parity-conserving transition (CPC). For the phase diagram, see Fig.10. On the basis of Eq. (3), the $`t`$-dependence of the magnetization in scaling form may be written as $$m(t,h)t^{\frac{\beta _s}{\nu Z}}\stackrel{~}{g}\left(ht^{\frac{\mathrm{\Delta }}{\nu Z}}\right)$$ (54) We have investigated the evolution of the nonequilibrium system from an almost perfectly magnetized initial state (or rather an ensemble of such states) . This state is prepared in such a way that a single up-spin is placed in the sea of down-spins at $`L/2`$. Using the language of kinks this corresponds to the usual initial state of two nearest neighbour kinks placed at the origin. At the critical point the density of spins is given as $$n_s(t,h)t^{\eta _s}g_1\left(ht^{\frac{\mathrm{\Delta }}{\nu Z}}\right)$$ (55) for the deviation of the spin density from its initial value, $`n_s=m(t,h)m\left(0\right)`$. The results are summarized in Table IV. Concerning static exponents, only the magnetic exponent $`\mathrm{\Delta }`$ remains unchanged under the effect of the critical fluctuations. This is not an obvious result (see, e.g., the coherence-length exponent above). $`\beta _s^{}`$, characterizing the level-off values of the survival probability of the spin clusters is defined through $$\underset{t\mathrm{}}{lim}P_s(t,h)h^{\beta _{s}^{}{}_{}{}^{}};$$ (56) this is a new static exponent connected with others via the scaling law $$\beta _s{}_{}{}^{}=\frac{\delta _s\nu Z}{\mathrm{\Delta }}$$ (57) which is satisfied by the exponents obtained numerically, within error. ## VI Damage-spreading investigations While damage spreading (DS) was introduced in biology it has become an interesting topic in physics as well . The main question is if damage introduced in a dynamical system survives or disappears. To investigate this the usual technique is to construct a pair of initial configurations that differ at a single site, and let them evolve with the same dynamics and external noise. This method has been found very useful for accurate measurements of dynamical exponents of equilibrium systems . In we investigated the DS properties of several 1d models exhibiting a PC-class phase transition. The order parameter that we measured in simulations is the Hamming distance between the replicas: $$D\left(t\right)=\frac{1}{L}\underset{i=1}{\overset{L}{}}\left|s_is_i^,\right|$$ (58) where $`s_i`$ and the replica $`s_i^,`$ may denote now spin or kink variables. If there is a phase transition point, the Hamming distance behaves in a power law manner at that point. In case of initial replicas differing at a single site (seed simulations) this looks like: $$D\left(t\right)t^\eta ,$$ (59) Similarly the survival probability of damage variables behaves as: $$P_s\left(t\right)t^\delta $$ (60) and the average mean square distance of damage spreading from the center scales as: $$R^2\left(t\right)t^z.$$ (61) In case of the NEKIM model we have investigated the DS on both the spin and kink levels. We found that the damage-spreading point coincides with the critical point, and that the kink-DS transition belongs to the PC class. The spin damage exhibits a discontinuous phase transition, with compact clusters and PC-like spreading exponents. The static exponents determined by finite-size scaling are consistent with those of spins of the NEKIM model at the PC transition point and the generalised hyperscaling law is satisfied. By inspecting our large data set for various models exhibiting multiple absorbing states, we are led to a final conjecture: BAWe dynamics and $`Z_2`$ symmetry of the absorbing states together form the necessary condition to have a PC-class transition. ## VII Summary The present paper has been devoted to reviewing most of the authors’ investigations of a nonequilibrium kinetic Ising model (NEKIM). NEKIM has proven to be a good testing ground for ideas, especially connected with dynamical scaling, in the field of nonequilibrium phase transitions. The field covered here belongs to the class of absorbing-state phase transitions which are usually treated on the level of particles, the best known examples being branching annihilating random walk models. The phase transition on the level of kinks in the Ising system belongs the the so-called parity-conserving universality class. The NEKIM model offers the possibility of revealing and clarifying features and properties which take place on the level of the underlying spin system as well. The introduction of a magnetic field in the Ising problem has also led to understand different features, e.g., the change of the universality class from PC to DP in the critical behavior at (and in the vicinity of) the phase transition of kinks. Investigations of various critical properties have been carried out mainly with the help of numerical simulations (time-dependent simulations, finite-size scaling and dynamic early-time MC methods have been applied). In addition, a generalized mean-field approximation was used, particularly in the neighborhood of a limiting situation of the phase diagram of NEKIM, where the usually second-order phase transition line ends at a MF-type first-order point. Some of the critical exponents, especially that of the order parameter, could be predicted to high accuracy in this way. While from the side of numerics a lot of information has accumulated over the last five years concerning the PC universality class, further investigations are needed at the microscopic level. Acknowledgements The authors would like to thank the Hungarian research fund OTKA ( Nos. T-23791, T-025286 and T-023552) for support during this study. One of us (N.M.) would like to acknowledge support by SFB341 of the Deutsche Forschungsgemeinschaft, during her stay in Köln, where this work was completed. G. Ó acknowledges support from Hungarian research fund Bólyai (No. BO/00142/99) as well. The simulations were performed partially on the FUJITSU AP-1000 and AP-3000 supercomputers and Aspex’s System-V parallel processing system (www.aspex.co.uk).
warning/0001/nlin0001016.html
ar5iv
text
# Modelling thermostatting, entropy currents and cross effects by dynamical systems ## I Introduction There is a recent interest in modelling transport processes by simple dynamical systems with chaotic dynamics. One class of models, actually inspired by non-equilibrium molecular dynamics (NEMD) simulations, describes systems driven by external fields with a spatially uniform dynamics subjected to periodic boundary conditions . Another approach concentrates on systems driven from the boundaries, which lead to steady states with sustained gradients of the thermodynamic fields . For a comparatively simple, but as far as their transport properties are concerned, generic class of dynamical systems, the multibakers, we show that both mechanisms of driving can simultaneously be worked out. This leads to an improved understanding of the relation between the approaches. In the former approach transport is driven by a field acting uniformly in the full system, while in the latter case the driving is concentrated to a microscopic region in space. From this point of view, boundary-driven transport is closely analogous to transport in a dynamical system with periodic boundary conditions, which is driven out of equilibrium by a “perturbation” of the dynamics, localised in a macroscopically small region. In all models for transport, as emphasized by Nicolis and coworkers , a quantity of central interest is the heat flux, or equivalently the entropy flux, from the system into its environment. A central aim of modelling transport by dynamical systems is to identify settings, which are consistent with the thermodynamic entropy balance $$\frac{\mathrm{d}S}{\mathrm{d}t}=\frac{\mathrm{d}_eS}{\mathrm{d}t}+\frac{\mathrm{d}_iS}{\mathrm{d}t},$$ (1) i.e., with the statement that the temporal change of the thermodynamic entropy $`S`$ can be decomposed into two contributions, called the external and internal change of the entropy, respectively. This integral form can be rewritten into a local balance equation when the two terms on the right hand side correspond to integrals of local densities. In that case, the time derivative of the entropy density $`s`$ appears as $$_ts=\mathrm{\Phi }+\sigma ^{(\mathrm{irr})}$$ (2) where $`\mathrm{\Phi }`$ and $`\sigma ^{(irr)}`$ represent the densities of the entropy flux and the rate of irreversible entropy production, respectively. In the bulk of typical macroscopic systems the entropy flux can be written as the divergence of the entropy current $`j^{(s)}`$, $$\mathrm{\Phi }=j^{(s)},$$ (3) reflecting the fact that no heat can be taken out from the system locally . On the other hand, this form has to be generalized at positions where there is a heat current flowing into an attached thermostat, and in cases where the entropy current is not differentiable, like for istance across interfaces between different materials. In those cases the entropy flux is not a full divergence, and it need not even be defined as a density. Rather the flux should then be written as $$\mathrm{\Phi }=j^{(s)}+\mathrm{\Phi }^{(th)},$$ (4) where $`j^{(s)}`$ is still the entropy current flowing in the system, but $`\mathrm{\Phi }^{(th)}`$ accounts for the heat taken out in the form of a direct flow into the surroundings, which acts then as a thermostat. In the present paper, we shall put special emphasis on the role of the entropy flux $`\mathrm{\Phi }^{(th)}`$, and on exploring under which conditions it can vanish. We find conditions on how to model the entropy balance for thermodynamic bulk systems, and for macroscopic systems, which are subjected to thermostatting by either a localized sink for the entropy or a spatially uniform coupling to a thermostat. The role of the thermodynamic entropy of dynamical systems is played by the coarse-grained Gibbs entropy, whose usefulness in understanding irreversibility from the point of view of dynamical systems is by now thoroughly explained in the literature . ( For stochastically perturbed dynamical systems where noise generates a kind of coarse graining, see ). The bulk dynamics is represented by a multibaker model driving two fields, the density $`\varrho `$ and the the kinetic energy per particle $`T`$, with a local source density $`q`$ for the latter . A connection with macroscopic transport equations is aimed at in a suitable defined continuum limit (the macroscopic limit), where the field $`T`$ will be interpreted as a temperature, based on the experience that this quantity is closely related to the average kinetic energy per particle. We shall consider a sequence of periodic models of increasing complexity. Model I corresponds to a homogeneous isothermal system described by a thermostatting algorithm. In this model no entropy current is defined — its entire entropy flux stems from a $`\mathrm{\Phi }^{(th)}`$. By allowing a spatial resolution of the isothermal system (Model II), a non-vanishing $`j^{(s)}`$ term appears in the transient behaviour, but $`\mathrm{\Phi }^{(th)}`$ remains unchanged. It is the only contribution to the flux in a steady state. Model III is still isothermal but with a locally deviating dynamics in one of the multibaker cells representing a boundary. The bulk dynamics can then be chosen so that (3) holds in the bulk, and all the heat taken out is concentrated in the boundary with a $`\mathrm{\Phi }^{(th)}0`$ there. In model IV we allow for temperature changes and local heat sources. By taking $`q`$ locally deviating from that of the bulk in one cell, we find a steady temperature profile with a break at the boundary. The $`q`$ distribution can then be chosen such that again (3) holds in the bulk. The heat source in the boundary is however singular. It corresponds exactly to the one which follows from thermodynamics. Finally, we consider a multibaker chain joined together from two subchains with different material properties. This models a junction between two metals so that one can observe thermodynamic cross effects, like the Peltier and Seebeck effects, very much in the same arrangements as in classical experiments. This paper is organized as follows. In Section II the local dynamics of the considered multibaker model is defined, and its local entropy balance is worked out. In Section III, Model I – Model IV are treated, which share periodic boundary conditions and represent thermodynamic settings of increasing complexity. Section IV is devoted to cross-effects. We conclude with a short discussion in Section V. ## II Local transport and thermodynamic relations for multibakers In this section we describe the local dynamics of a cell of a multibaker map modelling a system with particle and heat transport . We work out its density and kinetic-energy dynamics, and present general relations for the entropy changes. The effect of boundary conditions will be considered in subsequent sections for a few models with progressively richer thermodynamics. The phase space $`(x,p)`$ of the multibaker map consists of cells labelled by the index $`m`$ (Fig. 1). The division of the $`x`$ axis into cells corresponds to a partitioning of the configuration space into regions, sufficiently large to allow to characterize the state inside the cell by thermodynamic variables and small enough to neglect variation of these variables on the length scale of the cells (local equilibrium approximation). Every cell has a width $`a`$ and height $`b1`$. The coordinates of individual particles in the cell are given as a position variable $`x`$, and a momentum-like variable $`p`$. We are interested in the dynamics of two dimensionless fields, the phase-space density $`\varrho (x,p)`$, and a field $`T(x,p)`$ characterizing the local kinetic energy per particle. After each time unit $`\tau `$, every cell is divided into three columns (Fig. 1) with respective widths $`al_m`$, $`as_m`$ and $`ar_m`$. (Note that $`l_m+s_m+r_m=1`$ for any $`m`$.) The right (left) column of width $`ar_m`$ ($`al_m`$) is uniformly squeezed and stretched into a strip of width $`a`$ and of height $`l_{m+1}`$ ($`r_{m1}`$), which is mapped to the right (left) neighbouring cell. The middle one preserves its area and remains in cell $`m`$. Note that the map is one-to-one on its domain. It globally preserves the phase-space volume, but it can nevertheless locally expand or contract the phase-space volume. In Ref. it was argued that only the choice of the contraction factors given here can be consistent with thermodynamics (in fact, one can find an analogous formulation with a fully area preserving dynamics at the expense of a spatial variation of the volume of the cells of the multibaker; cf. ). The field $`T`$ is advected by the particle dynamics, and — in order to mimic a local heating of the system — it is also multiplied by a factor $`(1+\tau q)`$ depending on the averages characterizing the local currents and the thermodynamic state. By this a mean-field-like coupling of the motion of the particles in and around of a given cell is introduced. In general, the width of the columns may depend on the variables characterizing the thermodynamic state in the vicinity of the cell, so that they vary in time and space. This is indicated by the explicit dependence of the parameters on the cell index. Iteration of these rules defines the time evolution of the system. The $`(x,p)`$ dynamics generates ever refining structures in the distributions $`\varrho (x,p)`$ and $`T(x,p)`$. For simplicity, we take the fields initially constant in each cell: $`\varrho (x,p)=\varrho _m`$, $`T(x,p)=T_m`$ . ### A Dynamics of the particle density and the particle current After one step of iteration, the fields will be piecewise constant on the strips defined in Fig. 1. Due to the conservation of particles, the phase-space density takes the respective values $$\varrho _{m,r}^{}=\frac{r_{m1}}{l_m}\varrho _{m1},\varrho _{m,s}^{}=\varrho _m,\varrho _{m,l}^{}=\frac{l_{m+1}}{r_m}\varrho _{m+1}.$$ (5) (The prime will always indicate quantities evaluated after one time step.) The contraction factors $`r_{m1}/l_m`$ and $`l_{m+1}/r_m`$ contribute to the (weighted) local phase-space contraction $`\sigma _m`$ of cell $`m`$: $$\sigma _m=\frac{1}{\tau }\left[\varrho _{m1}r_{m1}\mathrm{ln}\frac{r_{m1}}{l_m}+\varrho _{m+1}l_{m+1}\mathrm{ln}\frac{l_{m+1}}{r_m}\right].$$ (6) After one time step, the average density $`\varrho _m^{}`$ in cell $`m`$ is determined by its initial density $`\varrho _m`$ and by the initial densities $`\varrho _{m\pm 1}`$ of the neighbouring cells. Multiplying the strip densities (5) with the widths of the respective strips, adding them up and dividing the sum by the width $`a`$ of the cell, one obtains the average (or the coarse-grained) density after the iteration $$\varrho _m^{}=s_m\varrho _m+r_{m1}\varrho _{m1}+l_{m+1}\varrho _{m+1}.$$ (7) The coarse-grained density evolves according to this master equation, which can be rearranged to obtain the discrete conservation law of the density $$\frac{\varrho _m^{}\varrho _m}{\tau }=\frac{j_mj_{m1}}{a}.$$ (8) Here $$j_m=\frac{a}{\tau }(r_m\varrho _ml_{m+1}\varrho _{m+1})$$ (9) is the discrete particle current flowing through the right boundary of cell $`m`$. ### B The kinetic-energy dynamics and the energy current According to the $`T`$ dynamics described above, the updated values $`T_{m,r}^{}`$, $`T_{m,s}^{}`$, $`T_{m,l}^{}`$ for $`T`$ on the respective strips $`R`$, $`S`$, $`L`$ of cell $`m`$ contain a source term characterized by a local strength $`q_m`$: $`T_{m,r}^{}=T_{m1}\left[1+\tau q_m\right],`$ (10) $`T_{m,s}^{}=T_m\left[1+\tau q_m\right],`$ (11) $`T_{m,l}^{}=T_{m+1}\left[1+\tau q_m\right].`$ (12) This strength is yet undetermined. It depends on the physical setting of thermostatting to be modelled and on the average of $`\varrho `$ and $`T`$ values in the cells and in its neighbours. The $`(x,p)`$ dynamics also drives the field $`T`$, i.e., after one iteration the kinetic-energy density of cell $`m`$ takes the value $$\varrho _m^{}T_m^{}=[s_m\varrho _mT_m+r_{m1}\varrho _{m1}T_{m1}+l_{m+1}\varrho _{m+1}T_{m+1}](1+\tau q_m).$$ (13) This equation can be rearranged as a discrete balance equation for the time evolution of $`\varrho T`$: $$\frac{\varrho _m^{^{}}T_m^{^{}}\varrho _mT_m}{\tau }=\varrho _m^{^{}}T_m^{^{}}\frac{q_m}{1+\tau q_m}\frac{j_m^{(\varrho T)}j_{m1}^{(\varrho T)}}{a},$$ (14) where $`j_m^{(\varrho T)}=T_mj_m(a^2l_{m+1}/\tau )\varrho _{m+1}(T_{m+1}T_m)/a`$ is a corresponding discrete energy current. Note that the r.h.s of (14) is not a full divergence, in accordance with the fact that the kinetic energy is not a conserved quantity. In an isothermal system where there is no kinetic energy dynamics, no source can be present ($`q_m=0`$). ### C Gibbs entropy and the coarse-grained entropy In this study we are interested in both the temporal evolution of the exact fields $`\varrho (x,p)`$ and $`T(x,p)`$, and in the evolution of their respective cell averages $`\varrho _m`$ and $`T_m`$. The former densities characterize the microscopic time evolution, while the averages describe the local thermodynamic state in spatially small regions. Both levels of description admit entropy functionals, which are commonly denoted as Gibbs and coarse-grained entropy. The Gibbs entropy $`S^{(G)}`$ is related to the detailed knowledge of the system. It is taken with respect to the exact densities $`\varrho (x,p)`$ and $`T(x,p)`$. In a given cell it is defined as $$S_m^{(G)}=_{\text{over cell }m}dxdp\varrho (x,p)\mathrm{ln}\left(\frac{\varrho (x,p)}{\varrho ^{}}T(x,p)^\gamma \right).$$ (15) Here $`\varrho ^{}T^\gamma `$ plays the role of a local $`T`$-dependent reference density with a constant reference density $`\varrho ^{}`$ and $`\gamma `$ an as yet undetermined constant. The coarse-grained entropy $`S_m`$ has a similar form, but it is based on the averaged values in the considered cell: $$S_m=a\varrho _m\mathrm{ln}\left(\frac{\varrho _m}{\varrho ^{}}T_m^\gamma \right).$$ (16) As mentioned above, throughout the paper we only consider initial distributions, which are uniform in every cell (cf. for more general choices). As a consequence, initially $`S_m=S_m^{(G)}`$, and after one time step the entropies become $$S_m^{(G)^{}}=a\left[s_m\varrho _m\mathrm{ln}\left(\frac{\varrho _m}{\varrho ^{}}T_{m,s}^{}_{}{}^{}\gamma \right)+r_{m1}\varrho _{m1}\mathrm{ln}\left(\frac{\varrho _{m,r}^{}}{\varrho ^{}}T_{m,r}^{}_{}{}^{}\gamma \right)+l_{m+1}\varrho _{m+1}\mathrm{ln}\left(\frac{\varrho _{m,l}^{}}{\varrho ^{}}T_{m,l}^{}_{}{}^{}\gamma \right)\right],$$ (17) and $$S_m^{}=a\varrho _m^{}\mathrm{ln}\left(\frac{\varrho _m^{}}{\varrho ^{}}T_m^{}_{}{}^{}\gamma \right).$$ (18) ### D Entropy balance The coarse-grained entropy fulfills a local entropy balance in direct analogy to the one in irreversible thermodynamics. To derive this equation one identifies at any given time the difference $`S_mS_m^{(G)}`$ as the information on the microscopic state of the system, which cannot be resolved in the coarse-grained description. The temporal change of this lack of information is then identified with the irreversible entropy production $`\mathrm{\Delta }_iS_m`$, and the change $`(S_m^{(G)}S_m^{(G)})`$ of the Gibbs entropy with the entropy flux $`\mathrm{\Delta }_eS_m`$. Thus, $$\frac{S_m^{}S_m}{\tau }=\frac{\mathrm{\Delta }_eS_m}{\tau }+\frac{\mathrm{\Delta }_iS_m}{\tau }$$ (19) which is a discrete analog of (1). The form of the entropy production is \[cf. (17) and (18)\]: $`{\displaystyle \frac{\mathrm{\Delta }_iS_m}{\tau }}`$ $`=`$ $`{\displaystyle \frac{[S_m^{}S_m^{(G)^{}}][S_mS_m^{(G)}]}{\tau }}`$ (20) $`=`$ $`{\displaystyle \frac{a}{\tau }}\left[\varrho _m^{}\mathrm{ln}\left({\displaystyle \frac{\varrho _m^{}T_m^\gamma }{\varrho _mT_{m,s}^\gamma }}\right)+\varrho _{m1}r_{m1}\mathrm{ln}\left({\displaystyle \frac{\varrho _{m,r}^{}T_{m,r}^\gamma }{\varrho _mT_{m,s}^\gamma }}\right)+\varrho _{m+1}l_{m+1}\mathrm{ln}\left({\displaystyle \frac{\varrho _{m,l}^{}T_{m,l}^\gamma }{\varrho _mT_{m,s}^\gamma }}\right)\right],`$ (21) where we used that $`S_mS_m^{(G)}`$ vanishes due to the particular choice of initial conditions. The entropy flux becomes $$\frac{\mathrm{\Delta }_eS_m}{\tau }=\frac{a}{\tau }\left[(\varrho _m^{^{}}\varrho _m)\mathrm{ln}\left(\frac{\varrho _m}{\varrho ^{}}T_m^\gamma \right)+\varrho _m^{^{}}\mathrm{ln}\frac{T_{m,s}^\gamma }{T_m^\gamma }+\varrho _{m1}r_{m1}\mathrm{ln}\left(\frac{\varrho _{m,r}^{}}{\varrho _m}\frac{T_{m,r}^\gamma }{T_{m,s}^\gamma }\right)+\varrho _{m+1}l_{m+1}\mathrm{ln}\left(\frac{\varrho _{m,l}^{}}{\varrho _m}\frac{T_{m,l}^{}{}_{}{}^{\gamma }}{T_{m,s}^{}{}_{}{}^{\gamma }}\right)\right]$$ (22) which can be split into a divergence of an entropy current and a flux into the thermostat $$\frac{\mathrm{\Delta }_eS_m}{a\tau }=\frac{j_m^{(s)}j_{m1}^{(s)}}{a}+\mathrm{\Phi }_m^{(\mathrm{th})}$$ (23) with $`j_m^{(s)}`$ $``$ $`j_m\mathrm{ln}\left({\displaystyle \frac{\varrho _m}{\varrho ^{}}}T_m^\gamma \right)+{\displaystyle \frac{al_{m+1}}{\tau }}\varrho _{m+1}\mathrm{ln}\left({\displaystyle \frac{\varrho _{m+1}}{\varrho _m}}{\displaystyle \frac{T_{m+1}^\gamma }{T_m^\gamma }}\right)\varrho _mv_m,`$ (25) $`\mathrm{\Phi }_m^{(\mathrm{th})}`$ $``$ $`{\displaystyle \frac{1}{\tau }}\left[\varrho _m^{}\mathrm{ln}{\displaystyle \frac{T_{m,s}^{}_{}{}^{}\gamma }{T_m^\gamma }}+r_{m1}\varrho _{m1}\mathrm{ln}\left({\displaystyle \frac{r_{m1}}{l_m}}{\displaystyle \frac{T_{m,r}^{}_{}{}^{}\gamma T_m^\gamma }{T_{m,s}^{}_{}{}^{}\gamma T_{m1}^\gamma }}\right)+l_{m+1}\varrho _{m+1}\mathrm{ln}\left({\displaystyle \frac{l_{m+1}}{r_m}}{\displaystyle \frac{T_{m,l}^{}_{}{}^{}\gamma T_m^\gamma }{T_{m,s}^{}_{}{}^{}\gamma T_{m+1}^\gamma }}\right)\right]`$ (26) $``$ $`{\displaystyle \frac{v_m\varrho _mv_{m1}\varrho _{m1}}{a}}.`$ (27) Note that (23) is a discrete counterpart of (4), and $`j_m^{(s)}`$ and $`\mathrm{\Phi }_m^{(th)}`$ are the discrete entropy current and entropy flux to the thermostat, respectively. ### E The macroscopic limit The projection of the multibaker dynamics on the $`x`$ axis corresponds to a biased random walk with some diffusion coefficient and drift. The drift has to be present if we want to model nonequilibrium systems subjected to electric fields and/or temperature gradients. The requirement of consistency with an advection diffusion equation in the large system and long time limit, when the cell size is much smaller than the system size, and the time unit is much shorter than the macroscopic relaxation time, leads to the scaling relation: $`r_m`$ $`=`$ $`{\displaystyle \frac{\tau D}{a^2}}\left(1+{\displaystyle \frac{av_m}{2D}}\right),`$ (29) $`l_m`$ $`=`$ $`{\displaystyle \frac{\tau D}{a^2}}\left(1{\displaystyle \frac{av_m}{2D}}\right).`$ (30) Here we allow for a location dependence of the drift $`v_m`$ but assume the diffusion coefficient $`D`$ to be spatially constant. The continuum limit of the multibaker dynamics, which is taken with these constraints, is called the macroscopic limit. Formally, it corresponds to taking $`a,\tau 0`$ while keeping $`D`$ fixed, and $`am`$, $`\varrho _m`$, $`T_m`$, $`v_m`$ and $`q_m`$ finite so that they approach a macroscopic position coordinate $`x`$, and smooth functions $`\varrho (x)`$, $`T(x)`$, $`v(x)`$ and $`q(x)`$, respectively. After taking this limit we call $`\varrho (x)`$ the density and $`T(x)`$ temperature distribution in the system. The macroscopic limit of all the local relations given in (7)–(22) can be worked out explicitly . In the following this limit will be indicated by an arrow ‘$``$’ . Here we only mention that the system’s equation of state turns out to be that of classical ideal gas with $`\gamma \varrho `$ as its heat capacity (measured in units of Boltzmann’s constant). ## III Periodic models ### A Model I: Isothermal single-cell multibaker We start by discussing the simplest conceivable model for describing a macroscopic transport process. A particle current induced by an external field in an isothermal environment described by a single baker cell subjected to periodic boundary conditions. The right boundary of the cell is identified with its left boundary and the mapping is from the cell onto itself. Because of driving $`rl`$, and thermostatting is applied via the appearance of the contraction rates $`l/r`$ and $`r/l`$ in order to reach a steady state. This mapping propagates the coordinates of a large number of particles, which do not interact, i.e., they all are mapped by the same mapping. Clearly, this system does not admit a spatial resolution of the densities characterizing the transport process. Its local and global behaviour coincides, so that the subscript $`m`$ of the densities can be descarded in this case. For the single-cell multibaker the master equation (7) predicts $`\varrho ^{}=\varrho \overline{\varrho }.`$ This implies that the model is describes tranport in a steady state with the average density $`\overline{\varrho }`$. The particle curent $`j=(a/\tau )(rl)\overline{\varrho }=v\overline{\varrho }`$ is constant in space and time, and the entropy production (21) becomes ($`T=const.`$). $$\frac{\mathrm{\Delta }_iS}{a\tau }=\overline{\varrho }\frac{(rl)}{\tau }\mathrm{ln}\left(\frac{r}{l}\right)\sigma $$ (31) which has the macroscopic limit: $$\frac{\mathrm{\Delta }_iS}{a\tau }\sigma ^{(irr)}=\overline{\varrho }\frac{v^2}{D}.$$ (32) In a similar way, the entropy flux (23, 27) has the macroscopic form: $$\frac{\mathrm{\Delta }_eS}{a\tau }\mathrm{\Phi }^{(\mathrm{th})}=\overline{\varrho }\frac{v^2}{D}$$ (33) As expected in a steady state, $`\mathrm{\Delta }_iS`$ and $`\mathrm{\Delta }_eS`$ add up to zero. More interestingly, however, these contributions to the change of entropy are also directly proportional to the local phase-space contraction (6), which reduces to $`\overline{\varrho }v^2/D`$ in the macroscopic limit. ### B Model II: Isothermal multibaker chain for field driven transport In the single-cell multibaker the coarse-grained density cannot evolve in time, since no spatial variations are resolved. For this reason, one also cannot distuinguish between the local entropy balance as described by (2), (4) and the balance (1) for the full macroscopic system. In order to address these points, we generalize the previous setting by considering a multibaker chain of $`N+1`$ cells, with spatially constant driving (i.e., $`r_m=r`$ and $`l_m=l`$) and periodic boundary conditions: cell $`N+1`$ and cell $`0`$ are identified. From (9) we obtain for the particle current $$j_m=\frac{a}{\tau }\left[(rl)\varrho _ml(\varrho _{m+1}\varrho _m)\right]$$ (34) which has the macroscopic limit: $$j_mj\varrho vD_x\varrho .$$ (35) This current varies along the chain as long as the cell densitites evolve in time. The asymptotic state, however, is formed by a spatially uniform density distribution with the average density $`\overline{\varrho }`$. For constant temperature $`T`$ the irreversible entropy production (21) becomes $$\frac{\mathrm{\Delta }_iS_m}{a\tau }=\frac{1}{\tau }\left[\varrho _m^{}\mathrm{ln}\frac{\varrho _m^{}}{\varrho _m}+r\varrho _{m1}\mathrm{ln}\left(\frac{\varrho _{m1}}{\varrho _m}\frac{r}{l}\right)+l\varrho _{m+1}\mathrm{ln}\left(\frac{\varrho _{m+1}}{\varrho _m}\frac{l}{r}\right)\right].$$ (36) Using (7) and (II E), a lengthy but straightforward calculation for $`a,\tau 0`$ shows that this local form is consistent with thermodynamics since in a general non-steady state it approaches $$\sigma ^{(irr)}=\frac{(\varrho vD_x\varrho )^2}{\varrho D}=\frac{j^2}{\varrho D}$$ (37) in the macroscopic limit. In a similar way , the entropy flux (22) can be calculated, which in the macroscopic limit has the form (4), with the entropy current $$j^{(s)}=j\left[1+\mathrm{ln}(\varrho /\varrho ^{})\right]$$ (38) and the entropy flux $$\mathrm{\Phi }^{(th)}=\frac{vj}{D}$$ (39) transfered directly to the environment. This local expression expresses that every cell is coupled to the thermostat. Due to the additional spatial resolution (as compared to Model I) the local and global features of the entropy balance can be different. It is worth considering the global entropy production, defined as \[cf. (36)\] $$\underset{m=0}{\overset{N}{}}\frac{\mathrm{\Delta }_iS_m}{\tau }=\frac{a}{\tau }\left[\underset{m=0}{\overset{N}{}}\varrho _m\right](rl)\mathrm{ln}\left(\frac{r}{l}\right)+\frac{a}{\tau }\underset{m=0}{\overset{N}{}}[\varrho _m^{}\mathrm{ln}\varrho _m^{}+\varrho _m\mathrm{ln}\varrho _m].$$ (40) Here, the logarithm of the ratios of densities of neighbouring cells in (36) drop out in the sum due to the periodic boundary conditions. The global entropy production then takes the macroscopic form $$\underset{m=0}{\overset{N}{}}\frac{\mathrm{\Delta }_iS_m}{\tau }\frac{d_iS}{dt}=\left[\underset{m=0}{\overset{N}{}}a\varrho _m\right]\frac{v^2}{D}\underset{m=0}{\overset{N}{}}a_t[\varrho \mathrm{ln}\varrho ]=𝒩\frac{v^2}{D}+\frac{dS}{dt},$$ (41) where $`𝒩=a\overline{\varrho }(N+1)`$ is the total number of particles in the multibaker chain, and $`S=_mS_m`$ is the total entropy. The global form of the entropy flux is \[cf. (22)\] $$\underset{m=0}{\overset{N}{}}\frac{\mathrm{\Delta }_eS_m}{\tau }=\frac{a}{\tau }\left[\underset{m=0}{\overset{N}{}}\varrho _m\right](rl)\mathrm{ln}\left(\frac{r}{l}\right)=\underset{m=0}{\overset{N}{}}a\sigma _m,$$ (42) which takes the macroscopic limit $$\underset{m}{}\frac{\mathrm{\Delta }_eS_m}{\tau }\frac{d_eS}{dt}=𝒩\frac{v^2}{D}.$$ (43) This shows that the local and the global entropy balances are markedly different. Locally, the entropy current depends only on the local current and the local density. There is only an indirect influence of the drift velocity $`v`$ through its contribution to the current and its influence on the density profile. In contrast, the macroscopic flux only depends on the total number of particles in the system and on the drift velocity. It is constant in time since it neither depends on the current nor on the density profile, which in general both evolve in time. Interestingly, also in this more general setting the negative of the global entropy flux equals the (total) phase space contraction at any time. This is in full harmony with the result obtained for the entropy flux in noisy dynamical systems by Nicolis and Daems . In contrast, the total irreversible entropy production is in general no longer directly related to the phase-space contraction. Its local form exactly amounts to Joule’s heating $`j^2/\varrho D`$, and globally it picks up an additional contribution characterizing the time-evolution of the macroscopic states. Therefore, the often cited relation between the global entropy production and the phase-space contraction only holds in steady states, where the modulus of the entropy flux coincides with the rate of entropy production. ### C Model III: Isothermal multibaker chain with driving localized to a single cell A generalization of the previous case is a system modelling boundary-driven transport. Here, we consider a slightly different setting with periodic boundary conditions. Away from the point of driving the macroscopic properties correspond to that of boundary driven transport, but due to the periodic boundary conditions the model has a simpler structure as a dynamical system, which makes the analogy with our previous models more transparent. We consider a multibaker chain where cell $`0`$ has a behaviour different from the rest of the chain in the sense that its drift velocity is different from that of the bulk where $`r=l`$, (i.e., $`v=0`$) but the diffusion coefficient is the same. This implies $`r_0rll_00`$ (cf. (II E)). The driving in cell zero has the tendeny to generate an accumulation of particles right to it and a slowly decreasing density distribution on its left (we assume $`v_0>0`$, $`r_0>r`$). In a steady state, this leads to a linear density profile $$\varrho _m=\overline{\varrho }+\left(\frac{N1}{2}m+1\right)\delta \varrho \text{for }m=1\mathrm{}N\text{,}$$ (44) whose increment $`\delta \varrho `$ is uniquely determined by $`r_0`$ (or $`v_0`$). In cell zero we find $`\varrho _0=\overline{\varrho }`$. A substitution into the master equation for $`m=1`$ or $`N`$ leads to $$\delta \varrho =\overline{\varrho }\frac{2}{N+1}(\frac{r_0}{r}1).$$ (45) Again the average density $`\overline{\varrho }`$ is related to the number of particles $`𝒩`$ in the system via $`a\overline{\varrho }(N+1)=𝒩`$. By taking into account that according to (II E) $`r_0=r[1+av_0/(2D)]`$, we obtain that $`\delta \varrho `$ is indeed proportional to $`v_0`$: $$v_0=D\frac{\delta \varrho }{a}\frac{N+1}{\overline{\varrho }}=D\frac{\delta \varrho }{a}\frac{𝒩}{a\overline{\varrho }^2}.$$ (46) The steady state current is: $$j=\overline{\varrho }\frac{v_0}{2}D\frac{\varrho _1\overline{\varrho }}{a}=D\frac{\delta \varrho }{a}=\frac{v_0\overline{\varrho }}{N+1}.$$ (47) The diffusion current of cells $`N`$ and $`0`$ is much stronger than otherwise, due to the sudden jump in the densities, but in the steady state the surplus is exactly compensated by the local drift currents. Thus the particle transport in the steady states of Model II and Model III are equivalent provided the drift $`v`$ in the former coincides with $`D\delta \varrho /(a\overline{\varrho })`$ in the latter. As far as the steady state transport is concerned, it does not matter, whether one applies a small uniform field leading to a spatially uniform drift $`v`$ or a large one $`v_0=(N+1)v`$ in a single cell only. All thermodynamic relations of relevance can be worked out not only for the steady state (44), but for general non-steady states. The local forms of the entropy production and the entropy flux in the bulk are special cases of (37)-(39). Since $`r=l`$ in the bulk, there is no entropy flux flowing into the thermostat $`\mathrm{\Phi }^{(th)}=0`$ for $`m=2,\mathrm{}N1`$. On the other hand, the entropy flux in cell $`0`$ and its neighbours does contain a part which cannot be written as a divergence. Due to (27), the average density of the flux flowing into the thermostat turns out to be: $`\mathrm{\Phi }^{(\mathrm{th})}\mathrm{\Phi }_N^{(\mathrm{th})}+\mathrm{\Phi }_0^{(\mathrm{th})}+\mathrm{\Phi }_1^{(\mathrm{th})}`$ $`=`$ $`{\displaystyle \frac{1}{\tau }}\left[(r\varrho _1r_0\varrho _0)\mathrm{ln}\left({\displaystyle \frac{r_0}{r}}\right)(r\varrho _Nl_0\varrho _0)\mathrm{ln}\left({\displaystyle \frac{r}{l_0}}\right)\right]`$ (48) $``$ $`{\displaystyle \frac{v_0}{2}}{\displaystyle \frac{\varrho _1\varrho _N}{a}}\varrho _0{\displaystyle \frac{v_0^2}{4D}}{\displaystyle \frac{\varrho _1+\varrho _N}{2}}{\displaystyle \frac{v_0^2}{4D}}.`$ (49) In the last approximation we have assumed that $`av_02D`$. In the steady state $`\varrho _N\varrho _1=(N1)\delta \varrho =(N1)aj/D`$, $`\varrho _0=\frac{(\varrho _1+\varrho _N)}{2}=\overline{\varrho }`$, and thus $$\mathrm{\Phi }^{(th)}\frac{v_0j}{D}.$$ (50) In the macroscopic limit the quantity $`\delta \varrho /a`$ approaches the gradient $`_x\varrho `$ in the bulk, $`\overline{\varrho }`$ and $`𝒩`$ stay constant and thus $`v_0`$ in (46) is proportional to $`1/a`$. The driving is singularly strong and so is the entropy flux density into the thermostat. By integrating, however, over the volume of cell zero and its neighbours we obtain the total entropy flux into the thermostat $`av_0j/D=j^2𝒩/\overline{\varrho }^2D`$. It coincides with the macroscopic limit of the total entropy flux $`_{m=0}^N\mathrm{\Delta }_eS_m/\tau `$ since the integral of $`_xj^{(s)}`$ vanishes in a periodic system, i.e., $$\frac{d_eS}{dt}=\frac{j^2}{\overline{\varrho }D}\frac{𝒩}{\overline{\varrho }}$$ (51) This result is equivalent to the steady state version of (43) expressed by the current. In this model there is no need for taking out heat along the bulk, thermostatting is active in cell $`0`$ and its neighbours only. It extracts exactly the same entropy flux there as the full entropy flux of Model II in the steady state . Thus the models refer to two different realizations of thermostatting the transport process. Model II should be viewed as e.g. a wire which is kept at constant temperature by removing the heat due to dissipation everywhere along its length — Model III is closer related to a thermally isolated system, where heat is tranported to the “boundary”, from where the system is driven. For the multibaker this takes place in the special cell $`m=0`$. Boundary driven transport typically leads, however, to non-uniform temperature profiles. A full treatment of such transport processes should be based therefore on a multibaker chain with kinetic energy dynamics. ### D Model IV: Multibaker chain with thermostatting localized to a single cell In the realm of classical thermodynamics heat is transported to the boundaries of the system. If, however, there is a break in the temperature profile, a jump in the heat current occurs and heat is taken out at this point. In order to model such a situation, we consider a multibaker chain with kinetic-energy dynamics. The general relation (21) and a calculation similar to the one leading to (37) yields in the macroscopic limit for the entropy production $$\sigma ^{(irr)}=\lambda \left(\frac{_xT}{T}\right)^2+\frac{j^2}{\varrho D}$$ (52) where $`j=\varrho vD_x\varrho `$ is the particle current, and $`\lambda =\gamma \varrho D`$ is the heat conductivity of the model. From (22)-(27) we obtain $$\mathrm{\Phi }^{(th)}=\gamma \varrho q\frac{vj}{D}$$ (53) as the entropy flux let directly into the thermostat, and $$j^{(s)}=\lambda \frac{_xT}{T}+\frac{e\mathrm{\Pi }}{T}j,$$ (54) as the entropy current with $`\mathrm{\Pi }`$ the bulk Peltier coefficient $$\frac{e\mathrm{\Pi }}{T}=\left(1+\mathrm{ln}\frac{\varrho T^\gamma }{\varrho ^{}}\right).$$ (55) Note that there is always a possibility to ’close’ the system locally in the sense that the source term $`q=q^{}=vj/\lambda `$ is chosen such that $`\mathrm{\Phi }^{(th)}`$ vanishes. We consider a periodic chain with fixed transition probabilities ($`r_m=r`$, $`l_m=l`$, $`m=0.1,..,N`$). The local heating sources are assumed to be constant in the bulk: $`q_m=q^{}`$ for $`m=1,\mathrm{},N`$ which differs from the source $`q_0`$ of cell $`0`$. In the steady state we find then a constant particle density along the chain. Inside the bulk, the kinetic-energy equation (13) implies for the steady temperature distribution $$T_m=[(1rl)T_m+rT_{m1}+lT_{m+1}](1+\tau q^{}).$$ (56) With periodic boundary conitions ($`T_{m=0}=T_0`$, and $`T_{m=N+1}=T_0`$) this equation has the following general solution: $$T_m=\frac{T_0}{\mathrm{sin}[b(N+1)]}\left(\frac{r}{l}\right)^{\frac{m}{2}}\left\{\mathrm{sin}[b(N+1)bm]+\left(\frac{l}{r}\right)^{\frac{N+1}{2}}\mathrm{sin}(bm)\right\}$$ (57) where $$\mathrm{cos}b=\sqrt{rl}\left(1\frac{\tau q^{}}{(1+\tau q^{})(r+l)}\right).$$ (58) The solution (57) has a break in cell zero, in the sense that the left and right derivatives are different. Only at the end of the system is the entropy flux not a full divergence. Applying the kinetic-energy equation (13) to cell zero in a steady state, where the density is constant, we find that: $$q_0=\frac{1}{\tau }\frac{r(T_0T_N)+l(T_0T_1)}{(1rl)T_0+rT_N+lT_1}.$$ (59) In the macroscopic limit $$aq_0\frac{D}{T}\left[_xT|_{(0)}_xT|_{(+0)}\right].$$ (60) This implies that the source density $`q_0`$ is singular but the total source $`Q_0=aq_0`$ inside cell $`0`$, is finite. It is worth comparing this with the thermodynamic treatment of the same problem. If there is a jump in the entropy current, in order to have finite entropy flux density $`\mathrm{\Phi }`$ in each point, it is unavoidable to allow for a $`\stackrel{~}{\mathrm{\Phi }}`$ which is not a full divergence: $$\mathrm{\Phi }=_xj^{(s)}+\stackrel{~}{\mathrm{\Phi }}.$$ (61) The form of $`\stackrel{~}{\mathrm{\Phi }}`$ one obtains by integrating (61) around the point where the jump in the derivative appears ($`x=0`$): $$_ϵ^ϵ\mathrm{\Phi }𝑑x=j^{(s)}|_{(ϵ)}^{(+ϵ)}+_ϵ^ϵ\stackrel{~}{\mathrm{\Phi }}𝑑x.$$ (62) The smoothness of $`\mathrm{\Phi }`$ implies that for $`ϵ0`$ $$_ϵ^ϵ\stackrel{~}{\mathrm{\Phi }}𝑑x=j^{(s)}|_{(0)}^{(+0)}.$$ (63) The multibaker result (53) implies that if $`q`$ is singular as in cell 0, then $`\mathrm{\Phi }_0^{(th)}=\gamma \varrho q_0`$. We then immediately see that $`\mathrm{\Phi }_0^{(th)}`$ is the analog of $`\stackrel{~}{\mathrm{\Phi }}`$. Indeed, by inserting the expression (54) for $`j^{(s)}`$ we have: $$_ϵ^ϵ\stackrel{~}{\mathrm{\Phi }}𝑑x=\frac{\lambda }{T}[_xT|_{(+0)}_xT|_{(0)}]$$ (64) which, on account of $`\lambda =\gamma \varrho D`$, exactly corresponds to (59). We have shown, that the thermodynamic evaluation and the macroscopic limit of $`q_0`$ lead to the same result. Physically this means, that by a proper choice of the source terms even the singularity in the entropy flux can be described in full harmony with thermodynamics, and the flux let to flow in the thermostat is exactly the amount of heat what is taken out also in the thermodynamic description if a break appears. ## IV Cross effects in a multibaker model Thermodynamic cross effects, which probe the (off-diagonal) Onsager coefficients, are difficult to observe in homogenous systems. When two materials are put into contact, however, they play a dominant role in understanding the heat and entropy currents. In order to mimick such phenomena, we consider two multibaker chains containing the cells $`m=M^{}\mathrm{}1`$ and $`m=0\mathrm{}M`$, respectively, $`M`$, $`M^{}>>1`$ which are brought into contact at $`m=0`$ (cf. Fig. 6). Now, the parameters $`l^{}`$, $`s^{}`$, $`r^{}`$ and $`l^+`$, $`s^+`$, $`r^+`$ in the two parts are different, and for generality we will also assume that the constant reference densities $`\varrho ^{,\pm }`$ are different. These differences will represent the different thermodynamic and transport properties of the materials. The difference in $`r`$ and $`l`$ gives rise to different drifts (conductivities) and diffusion coefficients, and the one in the reference density might be thought of reflecting for instance a different effective mass of the electrons. As in the previous subsection, the dynamics of this multibaker chain drives a density and a kinetic-energy field. In order to simplify the structure of the steady state density profiles, we restrict to the case $`r^+/l^+=r^{}/l^{}`$. This choice is motivated by a physical interpretation of $`r/l`$. After all, the macroscopic limit of $`r/l`$ is $`1+av/D`$, and $`v/D`$ is proportional to the external (electric) field, such that the requirement expresses that the external field should be the same in both materials. In the remainder of this section we discuss the transport in this model in two different settings: (i) a constant (non-vanishing) particle current and constant temperature; (ii) vanishing particle current and an isolated system which is only thermostatted at the “junction” $`m=0`$ and at the two “leads” $`m=M^{}`$ and $`m=M`$, respectively. Setting (i) allows us to discuss the Peltier effect and setting (ii) is used for the Seebeck effect. Before turning to these specific settings, however, we discuss the steady state profile of the (particle) density in general One does not expect noticable gradients in the electron density in either material, so that we fix them to the constant values $`\varrho ^{}`$ and $`\varrho ^+`$, leading to the spatially uniform current (cf. (9)) $`j={\displaystyle \frac{a}{\tau }}(r^+l^+)\varrho ^+={\displaystyle \frac{a}{\tau }}(r^{}l^{})\varrho ^{}.`$ In order to have the same current also across the junction, one has to require $`l^{}\varrho ^{}=l^+\varrho ^+`$ in adition. Together with the fact that $`v/D`$ is fixed for the whole system, this implies that there is a constant amount of Joule’s heating $`vj/D`$ per unit length of the system, which either has to be transferred to a local thermostat (cf. (53)), or leads to a local heating, i.e., enforces non-uniform temperature profiles. ### A The Peltier effect The requirement of a constant temperature in the setting of the Peltier effect requires the use of a thermostatted dynamics, $`q=0`$. In that case the Joule heating is transferred to the thermostat. Away from the junction, this leads to the flux $`\mathrm{\Phi }^{(th)}=vj/D`$. In the entropy balance the difference in the materials shows up only in the entropy currents. In view of (54), they become different in the two parts of the multibaker $`j^{(s,\pm )}=j\left(1+\mathrm{ln}\left[{\displaystyle \frac{\varrho ^\pm }{\varrho ^\pm }}T^\gamma \right]\right)={\displaystyle \frac{e\mathrm{\Pi }^\pm }{T}}j`$ implying that at the junction (i.e., between cells $`m=1`$ and $`m=0`$) an additional heat flux, the Peltier heat, is directed to the thermostat. It is characterized by the difference of the entropy currents $`j^{(s,+)}j^{(s,)}=j\mathrm{ln}{\displaystyle \frac{\varrho ^+/\varrho ^+}{\varrho ^{}/\varrho ^{}}}=j\left[\mathrm{ln}{\displaystyle \frac{l^+}{l^{}}}+\mathrm{ln}{\displaystyle \frac{\varrho ^+}{\varrho ^{}}}\right]{\displaystyle \frac{e\mathrm{\Pi }^{(+/)}}{T}}j,`$ where $`\mathrm{\Pi }^{(+/)}`$ defines he mutual Peltier coefficient of the two materials. It characterizes the amount of Peltier heat produced per unit electric current, and is the difference of the material Peltier coefficients \[cf. (55)\] $$\mathrm{\Pi }^{(+/)}=\mathrm{\Pi }^+\mathrm{\Pi }^{}.$$ (65) as also found in thermodynamics. ### B The Seebeck effect The Seebeck effect is observed in a thermally isolated system, where the junction is kept at a temperature $`T_j`$ different from the temperature $`T_l`$ prescribed at the leads, i.e., for the multibaker we demand $`T_M^{}=T_M=T_l`$ and $`T_1=T_0=T_j`$. This setup corresponds to a non-uniform temperature field, and, due to this, also to gradients in the electro-chemical potential $`\mu `$. Because of the difference in the material properties, these gradients can add up to a net potential drop between the leads, even if both leads are kept at the same temperature and there is no particle current. This follows immediately from the formal definition of the particle current in its discrete version: $$j_m=\frac{\sigma _{el}}{e^2}\left[\frac{\mu _{m+1}\mu _m}{a}+e\alpha \frac{(T_{m+1}T_m)}{a}\right],$$ (66) where $`\sigma _{el}`$ is the conductivity, $`e`$ the electric charge, and $`\alpha `$ the Seebeck coefficient of the material. In the considered system, we then have for vanishing current $`\mu _M^{}\mu _M`$ $`=`$ $`\mu _M^{}\mu _1+\mu _1\mu _0+\mu _0\mu _M`$ (67) $``$ $`e\alpha ^{}(T_M^{}T_1)+\mu _1\mu _0e\alpha ^+(T_0T_M)=e(\alpha ^+\alpha ^{})(T_lT_j)+\mu _1\mu _0.`$ (68) Here we have assumed the Seebeck coefficients to be approximately constant in the two materials. The macroscopic limit implies $`\mu _1=\mu _0`$, and we obtain for the mutual Seebeck coefficient of the two materials $$\alpha ^{(+/)}\frac{\mu _M^{}\mu _M}{e(T_lT_j)}=\alpha ^+\alpha ^{}.$$ (69) It characterizes the strength of the potential drop $`\mu _M^{}\mu _M`$ between the leads induced by the temperature diffence $`T_lT_j`$ between the leads and the junction. An expression for $`\alpha ^{(+/)}`$ can be determined for the multibaker by rewriting the expression (35) for the current in the form (66). Taking immediately the macroscopic limit and observing that the electro-chemical potential can be split into a chemical part $`\mu _c`$ and a part $`e\varphi `$ due to the external electric field $`E_x\varphi `$, one obtains $`j={\displaystyle \frac{\sigma _{el}}{e^2}}\left[_x(\mu _c+e\varphi )+e\alpha _xT\right]={\displaystyle \frac{\sigma _{el}E}{e}}{\displaystyle \frac{\sigma _{el}}{e^2}}\left[_\varrho \mu _c_x\varrho +_T\mu _c_xT+e\alpha _xT\right]=v\varrho D_x\varrho .`$ (70) Here $`v`$ $`=`$ $`{\displaystyle \frac{\sigma _{el}E}{e\varrho }},`$ (72) $`D`$ $`=`$ $`{\displaystyle \frac{\sigma _{el}}{e^2}}_\varrho \mu _c,`$ (73) $`e\alpha `$ $`=`$ $`_T\mu _c.`$ (74) By the first two equations we recover well-known relations from thermodynamics . Eq. (74) provides us with a relation for the Seebeck coefficient. Since the equation of state of the “multibaker gas” is that of a classical ideal gas $`\mu _c^\pm =(\gamma +1)T+T\mathrm{ln}\left({\displaystyle \frac{\varrho ^\pm T^\gamma }{\varrho ^\pm }}\right),`$ one obtains $`e\alpha ^{(+/)}=e(\alpha ^+\alpha ^{})=e\mathrm{ln}{\displaystyle \frac{\varrho ^+T_l^\gamma }{\varrho ^+}}e\mathrm{ln}{\displaystyle \frac{\varrho ^{}T_l^\gamma }{\varrho ^{}}}=e\left[\mathrm{ln}{\displaystyle \frac{l^+}{l^{}}}+\mathrm{ln}{\displaystyle \frac{\varrho ^+}{\varrho ^{}}}\right]={\displaystyle \frac{e\mathrm{\Pi }^{(+/)}}{T_l}},`$ where (65) was used in the last step. This comparison expresses the validity of the Onsager relation $`\mathrm{\Pi }^{(+/)}=\alpha ^{(+/)}T`$ for this class of models. ## V Discussion In this paper we have described the local and global transport properties of multibakers with a density and an energy dynamics. This class of maps makes an analytical modelling of transport processes by a deterministic chaotic dynamics possible, and admits a macroscopic description consistent with various aspects of irreversible thermodynamics. The macroscopic description comprises the time evolution of the average density and the kinetic energy in small regions of the physical space (the cells of the multibaker). The former density is interpreted as the particle density, and the latter as a temperature field. The averages in the small regions are in the spirit of local thermodynamic equilibrium, and the continuum description of thermodynamics arises in a macroscopic limit where the spatial resolution of the tranport process is small compared to the system size (or any other relevant macroscopic length), and where a discrete time-scale used in the definition of the dynamics is much smaller than macroscopic time scales. The relevant concept of entropy for multibakers is the Gibbs entropy defined with respect to the average density in the cells normalized by a temperature-dependent reference density. It is called the coarse-grained entropy. Based on an information-theoretic interpretation of the entropy, a local entropy balance can be derived, which in the macroscopic limit can be fully consistent with irreversible thermodynamics. This agreement holds provided that (i) a particular choice of local phase-space contraction and expansion rates is incorporated in the time evolution of the density, which we identified as a time-reversible evolution of the mapping in previous work , (ii) the density in the entropy is normalized by a reference density with a power-law dependence on the average kinetic energy in the cell, and (iii) appropriate source terms are incorporated in the evolution equations of the kinetic-energy field. No meaningful macroscopic description can be found for multibakers with other choices of the phase-space contraction factors. Modification of the source terms leads to additional contributions in the local entropy balance, which are interpreted as local entropy fluxes into a thermostat. In particular, for vanishing source terms one can mimic a transport process in a system with a spatially uniform temperature, i.e., one obtains a setting reminiscent of NEMD simulations of transport processes. Once the connection between the deterministic dynamics of the multibaker and the corresponding local thermodynamic relations is established, one can apply it to discuss transport in different macroscopic settings. A number of models with periodic boundary conditions were discussed in order to shed light on the global entropy balance in such systems. We find that, up to a trivial factor, the average phase-space contraction amounts to the entropy flux to the environment. This supports an earlier heuristic argument of Ruelle and others , who connected the phase-space contraction to the irreversible entropy production in a steady state. In contrast to the claims of some of the latter authors (cf. for instance ) the connection between the irreversible entropy production and the phase-space contraction rate breaks down away from stationarity. In fact, the contraction rate is still connected to the entropy flux in that situation, but the flux is no longer related to the rate of irreversible entropy production. This was shown (a) for multibakers with a uniform thermostatting (Model II), i.e., for models reminiscent of NEMD algorithms, (b) for systems where the driving and thermostatting is applied in a macroscopically small region of the system (Model III), thus giving rise to sustained density gradients, and (c) systems with a uniform external field and localized thermostatting (Model IV). The former two models have constant temperature fields, while the latter one supports a temperature profile with a discontinuity in the first derivative at the positon of thermostatting. As expected from the existence of the local entropy balance, the results are fully consistent with the corresponding thermodynamic description of the transport process. They suggest an interesting conclusion on modelling transport in bulk systems by isothermal NEMD simulations: These methods are valid in an approximation where the considered volume is sufficiently small to neglect density and temperature gradients. In steady states, they are equivalent to models, where the currents are the same, but thermostatting is only applied at the boundaries of a macroscopic system. Since even state of the art simulations can hardly cope with more than $`10^9`$ particles, i.e., with integration volumes larger than about $`\mu `$m<sup>3</sup>, this approximation seems to be well-justified in numerical studies. On the other hand, this assumptions should be kept in mind when isothermal NEMD modelling is taken as basis of theoretical studies of transport processes (cf. for instance ). To further demonstrate the use of multibakers with density and energy fields, we also discussed thermoelectric cross effects. The description of the transport properties requires in that case information on the equation of state, since the Seebeck effect is defined in terms of differences of chemical potentials. In previous work it was shown that the classical ideal-gas equation holds for multibakers. This is meaningful since the time evolution of the multibaker can be considered as the one of particles with phase-space coordinates $`(x,p)`$, which only interact by a (weak) mean-field like coupling manifested in a dependence of the local parameters on the average densities. With this input the Peltier and Seebeck effect were modelled and the Onsager relation, connecting their respective transport coefficients, was derived. The validity of this relation for multibakers is not a trivial result. It heavily relies on the choices (i)–(iii) to find an entropy balance consistent with irreversible thermodynamics. Summarizing, we demonstrated that multibakers establish a straightforward modelling of various transport phenomena by deterministic, chaotic dynamics. They give insight in the general structure of such models by explicit analytical calculations. This was demonstrated by discussions of thermoelectric cross effects, and of the relation between the average phase-space contraction, entropy fluxes and the rate of irreversible entropy production. ###### Acknowledgements. This work is dedicated to Prof. Gregoire Nicolis on occasion of his 60th birthday. Our discussions at the Centre of Complex Systems and Nonlinear Phenomena helped us a lot, when we were starting this project in the summer of 1998. We are also grateful to J.R. Dorfman, B. Fogarassy, J. Hajdu, G. Tichy and H. Posch for enlightening discussions. The paper was completed during a joint stay at the Max-Planck Institute for the Physics of Complex Systems, which we gratefully acknowledge together with support from the Hungarian Science Foundation (OTKA T17493, T19483) and the TMR-network Spatially extended dynamics.
warning/0001/physics0001046.html
ar5iv
text
# Role of negative-energy states and Breit interaction in calculation of atomic parity-nonconserving amplitudes. ## Abstract It is demonstrated that Breit and negative-energy state contributions reduce the $`2.5\sigma `$ deviation \[S.C. Bennett and C.E. Wieman, Phys. Rev. Lett. 82, 2484 (1999)\] in the value of the weak charge of <sup>133</sup>Cs from the Standard Model prediction to 1.7$`\sigma `$. The corrections are obtained in the relativistic many-body perturbation theory by combining all-order Coulomb and second-order Breit contributions. The corrections to parity-nonconserving amplitudes amount to 0.6% in <sup>133</sup>Cs and 1.1% in <sup>223</sup>Fr. The relevant magnetic-dipole hyperfine structure constants are modified at the level of 0.3% in Cs, and 0.6% in Fr. Electric-dipole matrix elements are affected at 0.1% level in Cs and a few 0.1% in Fr. PACS: 31.30.Jv, 12.15.Ji, 11.30.Er Atomic parity-nonconserving (PNC) experiments combined with accurate atomic structure calculations provide constrains on “new physics” beyond the Standard Model of elementary-particle physics. Compared to high-energy experiments or low-energy $`ep`$ scattering experiments, atomic single-isotope PNC measurements are uniquely sensitive to new isovector heavy physics . Presently, the PNC effect in atoms was most precisely measured by Boulder group in $`{}_{}{}^{133}\mathrm{Cs}`$ . They determined ratio of PNC amplitude $`E_{\mathrm{PNC}}`$ to the tensor transition polarizability $`\beta `$ for $`7S_{1/2}6S_{1/2}`$ transition with a precision of 0.35%. In 1999, Bennett and Wieman accurately measured tensor transition polarizability $`\beta `$, and by combining the previous theoretical determinations of the $`E_{\mathrm{PNC}}`$ with their measurements, they have found a value of the weak charge for $`{}_{}{}^{133}\mathrm{Cs}`$ $`Q_\mathrm{W}=72.06(28)_{\mathrm{expt}}(34)_{\mathrm{theor}}`$ which differed from the prediction of the Standard Model $`Q_\mathrm{W}=73.20(13)`$ by 2.5 standard deviations. They also reevaluated the precision of the early 1990s atomic structure calculations , and argued that the uncertainty of the predicted $`E_{\mathrm{PNC}}`$ is 0.4%, rather than previously estimated 1%. This conclusion has been based on a much better agreement of calculated and recently accurately measured electric-dipole amplitudes for the resonant transitions in alkali-metal atoms. In view of the reduced uncertainty, the purpose of this Letter is to evaluate contributions from negative-energy states (NES) and Breit interaction. It will be demonstrated that these contributions correct theoretical $`E_{\mathrm{PNC}}`$ and the resultant value of the weak charge by 0.6% in Cs. It is worth noting, that due to the smallness of these contributions at what had been believed to be a 1% theoretical error in Cs, the previous calculations have either omitted , or estimated the contributions from Breit interaction only partially . The main focus of the previous ab initio calculations has been correlation contribution from the residual Coulomb interaction (i.e. beyond Dirac-Hartree-Fock level). In both calculations important chains of many-body diagrams have been summed to all orders of perturbation theory. This Letter also reports correction due to NES and Breit interaction for $`E_{\mathrm{PNC}}`$ in francium. The interest in Fr stems from the fact that analogous PNC amplitude is 18 times larger in heavier <sup>223</sup>Fr compared to Cs . The measurement of atomic PNC in Fr is pursued by Stony Brook group . The quality of theoretical atomic wave-functions at small radii is usually judged by comparing calculated and experimental hyperfine-structure magnetic-dipole constants $`A`$, and in the intermediate region by comparing electric-dipole matrix elements. It will be demonstrated that the corresponding corrections to all-order Coulomb values are at the level of a few 0.1%. The PNC amplitude of $`nS_{1/2}n^{}S_{1/2}`$ transition can be represented as a sum over intermediate states $`mP_{1/2}`$ $`E_{\mathrm{P}NC}={\displaystyle \underset{m}{}}{\displaystyle \frac{n^{}S|D|mP_{1/2}mP_{1/2}|H_W|nS}{E_{nS}E_{mP_{1/2}}}}`$ (1) $`+`$ $`{\displaystyle \underset{m}{}}{\displaystyle \frac{n^{}S|H_W|mP_{1/2}mP_{1/2}|D|nS}{E_{n^{}S}E_{mP_{1/2}}}}.`$ (2) The overwhelming contribution from parity-violating interactions arises from the Hamiltonian $$H_\mathrm{W}=\frac{G_F}{\sqrt{8}}Q_\mathrm{W}\rho _{\mathrm{n}uc}(r)\gamma _5,$$ (3) where $`G_F`$ is the Fermi constant,$`\gamma _5`$ is the Dirac matrix, and $`\rho _{\mathrm{n}uc}(r)`$ is the nuclear distribution. To be consistent with the previous calculations the $`\rho _{\mathrm{n}uc}(r)`$ is taken to be a Fermi distribution with the “skin depth” $`a=2.3/(4\mathrm{ln}3)`$ fm and the cutoff radius $`c=5.6743`$ fm for $`{}_{}{}^{133}\mathrm{Cs}`$ as in Ref. , and $`c=6.671`$ fm for <sup>223</sup>Fr as in Ref. . The PNC amplitude is customarily expressed in the units of $`10^{11}i(Q_\mathrm{W}/N)`$, where $`N`$ is the number of neutrons in the nucleus ($`N=78`$ for <sup>133</sup>Cs and $`N=136`$ for <sup>223</sup>Fr ). Atomic units are used throughout the Letter. The results of the calculations for $`{}_{}{}^{133}\mathrm{Cs}`$ are $`E_{\mathrm{P}NC}=0.905\times 10^{11}i(Q_\mathrm{W}/N)`$, Ref. and $`E_{\mathrm{P}NC}=0.908\times 10^{11}i(Q_\mathrm{W}/N)`$, Ref. . The former value includes a partial Breit contribution $`+0.002\times 10^{11}i(Q_\mathrm{W}/N)`$, and the latter does not. Both calculations are in a very close agreement if the Breit contribution is added to the value of Novosibirsk group. The reference many-body Coulomb value $$E_{\mathrm{P}NC}^C=0.9075\times 10^{11}i(Q_\mathrm{W}/N)$$ (4) is determined as an average of the two, with Breit contribution removed from the value of Notre Dame group. For <sup>223</sup>Fr $`E_{\mathrm{P}NC}=15.9\times 10^{11}i(Q_\mathrm{W}/N)`$, Ref. , this value does not include Breit interaction. Ab initio relativistic many-body calculations of wave-functions, like coupled-cluster type calculations of , to avoid the “continuum dissolution problem” , start from the no-pair Hamiltonian derived from QED . The no-pair Hamiltonian excludes virtual electron-positron pairs from the resulting correlated wave-function. If the no-pair wave-functions are further used to obtain many-body matrix elements, the negative-energy state (NES) contribution is missing already in the second order. Recently, it has been shown that the magnetic-dipole transition amplitude both in He-like ions and alkali-metal atoms can be strongly affected by the NES correction. The enhancement mechanism is due to vanishingly small lowest-order values and also due to mixing of large and small components of a Dirac wavefunction by magnetic-dipole operator. For the NES ($`E<mc^2`$) the meaning of large and small components is reversed, i.e., small component is much larger than large component. The mixing of large positive-energy component with small component of NES results in much larger one-particle matrix elements, than in the no-pair case. The $`2mc^2`$ energy denominators lessen the effect, but, for example, the Rb $`5S_{1/2}6S_{1/2}`$ magnetic-dipole rate is reduced by a factor of 8 from the no-pair value by the inclusion of NES . The inclusion of Breit interaction also becomes important and the size of the correction is comparable to the Coulomb contribution. Just as in the case of magnetic-dipole operator, the Dirac matrix $`\gamma _5`$ in the weak Hamiltonian Eq. (3) mixes large and small components of wavefunctions. Similar mixing occurs in the matrix element describing interaction of an electron with nuclear magnetic moment (hyperfine structure constant $`A`$). As demonstrated below, the relative effect for these operators is not as strong as in the magnetic-dipole transition case, since the lowest order matrix elements are nonzero in the nonrelativistic limit, but is still important. It is worth noting that the problem of NES does not appear explicitly in the Green’s function or mixed-parity approaches; however the correction due to Breit interaction still has to be addressed. The NES Coulomb corrections have to be taken into account explicitly in the “sum-over-states” method , employing all-order many-body values obtained with the no-pair Hamiltonian. The analysis is based on the $`V_{N1}`$ Dirac-Hartree-Fock potential wave-functions, with the valence and virtual orbitals $`m`$ calculated in the “frozen” potential of core orbitals $`a`$. The second-order correction to a matrix element of one-particle operator $`Z`$ between two valence states $`w`$ and $`v`$ is represented as $`Z_{wv}^{(2)}`$ $`=`$ $`{\displaystyle \underset{iv}{}}{\displaystyle \frac{z_{wi}b_{iv}}{ϵ_vϵ_i}}+{\displaystyle \underset{iw}{}}{\displaystyle \frac{b_{wi}z_{iv}}{ϵ_wϵ_i}}+`$ (5) $`+{\displaystyle \underset{ma}{}}`$ $`{\displaystyle \frac{z_{am}(\stackrel{~}{g}_{wmva}+\stackrel{~}{b}_{wmva})}{ϵ_a+ϵ_vϵ_mϵ_w}}+{\displaystyle \underset{ma}{}}{\displaystyle \frac{(\stackrel{~}{g}_{wavm}+\stackrel{~}{b}_{wavm})z_{ma}}{ϵ_a+ϵ_wϵ_mϵ_v}}.`$ (6) This expression takes into account the residual (two-body) Coulomb, $`g_{ijkl}`$, and two-body $`b_{ijkl}`$ and one-body $`b_{ij}=_a\stackrel{~}{b}_{iaja}`$ Breit interaction. Static form of Breit interaction is used in this work. The tilde denotes antisymmetric combination $`\stackrel{~}{b}_{ijkl}=b_{ijkl}b_{ijlk}`$. Subscript $`i`$ ranges over both core and excited states. Note that summation over states $`i`$ and $`m`$ includes negative-energy states. The NES correction to PNC amplitudes arises in two circumstances, directly from the sum in Eq. (2) and in the values of electric-dipole and weak interaction matrix elements. If the length-gauge of the electric-dipole operator is used, the direct contribution of NES in the amplitude Eq. (2) is a factor of $`10^{13}`$ smaller than the total amplitude, and will be disregarded in the following. The numerical summations are done using 100 positive- and 100 negative-energy wavefunctions in a B-spline representation obtained in a cavity with a radius of 75 a.u. The breakdown of second-order corrections to matrix elements of weak interaction for Cs is given in Table I. The all-order values from Ref. are also listed in the table to fix the relative phase of the contributions. The matrix elements are each modified at 0.6-0.7% level. Most of the correction arises from positive-energy Breit contribution, negative-energy states contribute at a smaller but comparable level. The contributions from NES due to one-body and two-body Breit interaction are almost equal ( $`B_{}^{(1)}B_{}^{(2)}`$) and, in addition, $`B_+^{(1)}2B_+^{(2)}`$. The same relations hold also in francium. The corrections to the relevant length-form matrix elements are overwhelmingly due to the one-body Breit interaction, and are at 0.1% level. For example, the all-order reduced matrix element $`6S_{1/2}D6P_{1/2}=4.478`$ Ref. is increased by 0.005, bringing the total 4.483 into an excellent agreement with experimental value 4.4890(65). Generally, the corrections reduce absolute values of the weak interaction matrix elements, and increase absolute values of the dipole matrix elements, therefore, their net contributions to $`E_{\mathrm{PNC}}`$ have an opposite sign. Matrix elements of weak interaction are affected more strongly, because of the sampling of wave-function in the nucleus, where relativity is important. As demonstrated in Ref. , the four lowest-energy valence $`mP_{1/2}`$ states contribute 98% of the sum in Eq. (2), and for the purposes of this work, limiting the sums to only these states is sufficient. The corrections to $`E_{\mathrm{PNC}}`$ are calculated first by replacing the weak interaction matrix elements with the relevant second-order contributions and at the same time using all-order dipole matrix elements, and second by taking all-order $`H_W`$ matrix elements, and replacing $`D`$ with the appropriate correction. In both cases the experimental energies are used in the denominators. The needed all-order matrix elements for Cs are tabulated in Ref. . The summary of corrections to $`E_{\mathrm{PNC}}`$ is presented in Table II. The modifications in the weak interaction matrix element provide a dominant correction. The contribution due to NES in the Coulomb part is insignificant, and is already effectively included in the reference many-body Coulomb value $`E_{\mathrm{P}NC}^C`$, Eq. (4). The reference value $`E_{\mathrm{P}NC}^C`$ is modified by the Breit contributions by 0.6%, almost two times larger than the uncertainty in the Boulder experiment . The modified value is $`E_{\mathrm{P}NC}^{C+B}(^{133}\mathrm{Cs})=0.902(36)\times 10^{11}i(Q_\mathrm{W}/N).`$A 0.4% uncertainty had been assigned to the above result following analysis of Bennett and Wieman . When $`E_{\mathrm{P}NC}^{C+B}`$ is combined with the experimental values of transition polarizability $`\beta `$ and $`E_{\mathrm{PNC}}/\beta `$ , one obtains for the weak charge $`Q_\mathrm{W}(^{133}\mathrm{Cs})=72.42(28)_{\mathrm{expt}}(34)_{\mathrm{theor}}.`$This value differs from the prediction of the Standard Model $`Q_\mathrm{W}=73.20(13)`$ by 1.7$`\sigma `$, versus 2.5$`\sigma `$ discussed in Ref. , where $`\sigma `$ is calculated by taking uncertainties in quadrature. The only previous calculation of Breit contribution to PNC amplitude in Cs has been performed by the Notre Dame group , using mixed-parity Dirac-Hartree-Fock formalism. The one-body Breit interaction has been included on equal footing with the DHF potential, but the linearized modification to one-body Breit potential due to $`H_\mathrm{W}`$ ($`V_{\mathrm{PNC}\mathrm{HFB}}`$ in notation of Ref. ) has been omitted. It is straightforward to demonstrate that because of this omission, the comparable contribution from two-body part of the Breit interaction has been disregarded. In units of $`10^{11}i(Q_\mathrm{W}/N)`$, the result of the present calculation for one-body Breit contribution is 0.003 versus 0.002 in Ref. . Such disagreement is most probably caused by different types of correlation contribution included in the two approaches. Treating one-body Breit together with the DHF potential effectively sums the many-body contributions from one-body Breit interaction to all orders, and presents the advantage of the scheme employed in Ref. . However, the dipole matrix elements and energies in the sum Eq. (2) are effectively included at the DHF level in the formulation of Ref. , in contrast to high-precision all-order values employed in the present work. The difference between the two values can be considered as a theoretical uncertainty in the value of the Breit correction. Clearly more work needs to be done to resolve the discrepancy. The accuracy of the present analysis can be improved if the one-body Breit interaction is embodied in DHF equations, and the many-body formulation starts from the resulting basis. However, to improve present second-order treatment of the two-body part of the Breit interaction, higher orders of perturbation theory have to be considered. Apparently the most important contribution would arise from terms linearized in the Breit interaction, i.e. diagrams containing one matrix elements of the Breit interaction and the rest of the residual Coulomb interaction. The Breit and NES corrections to PNC amplitude in heavier Fr are more pronounced. The <sup>223</sup>Fr PNC amplitude 15.9 $`\times 10^{11}i(Q_\mathrm{W}/N)`$ from Ref. is reduced by 1.1%. Using all-order dipole matrix elements from Ref. , the following corrections due to modifications in the $`h_W`$ are found (in units of $`10^{11}i(Q_\mathrm{W}/N)`$): one-body Breit $`B_\pm ^{(1)}=0.131`$, two-body Breit $`B_\pm ^{(2)}=0.053`$, and the $`C_{}`$ correction, implicitly included in Ref. , is $`0.003`$. As in the case of Cs, the all-order no-pair Coulomb result for reduced matrix element $`7P_{1/2}D7S_{1/2}=4.256`$ is increased by inclusion of the Breit interaction and NES by $`0.0011`$, a 0.3% modification, leading to a much better agreement with experimental value 4.277(8) . The modification of the $`E_{\mathrm{P}NC}`$ due to corrections in the dipole matrix elements is much smaller than in the case of $`h_\mathrm{W}`$. At present there is no tabulation of accurate matrix elements of weak interaction for Fr, and the influence on $`E_{\mathrm{PNC}}`$ due to the Breit contribution in dipole matrix elements is estimated from average of the modification of individual dipole matrix elements 0.2%. The net result decreases the reference Coulomb value for <sup>223</sup>Fr by 0.18 $`\times 10^{11}i(Q_\mathrm{W}/N)`$, and the corrected value is $`E_{\mathrm{P}NC}^{C+B}(^{223}\mathrm{Fr})=15.7\times 10^{11}i(Q_\mathrm{W}/N).`$ Finally, it is worth discussing Breit and NES contributions to hyperfine-structure magnetic-dipole constants $`A`$ for the states involved into PNC calculations. The all-order no-pair Coulomb values in the recent work have been corrected using a similar second-order formulation; no details of the calculation have been given. The explicit contributions listed in Table III will be useful for correcting ab initio many-body Coulomb values. The table presents the contributions for two lowest valence $`S_{1/2}`$ and $`P_{1/2}`$ states. The calculations are performed using a model of uniformly magnetized nucleus with a magnetization radius $`R_m`$ given in the table. One finds that the additional terms reduce values calculated in the no-pair Coulomb-correlated approach. For Cs the corrections are of order 0.2% for $`6S_{1/2}`$, 0.1% for $`7S_{1/2}`$, and 0.3% for $`6P_{1/2}`$ and $`7P_{1/2}`$. The relative contributions to hyperfine constants in heavier Fr are larger, accounting for 0.5% of the total value for $`7S_{1/2}`$, 0.4% for $`8S_{1/2}`$, and 0.6% for $`7P_{1/2}`$ and $`8P_{1/2}`$. This work demonstrates that the Breit and NES contributions are comparable to the remainder of Coulomb correlation corrections unaccounted for in modern relativistic all-order many-body calculations and hence have to be systematically taken into account. In particular, the Breit interaction contributes 0.6% to parity-nonconserving amplitudes in Cs and 1.1% in Fr. The correction for Cs is almost twice the experimental uncertainty and reduces the recently determined 2.5 $`\sigma `$ deviation in the value of weak charge from the Standard Model prediction to 1.7$`\sigma `$. Both hyperfine constants and electric-dipole matrix elements are affected at a few 0.1%. By including NES and Breit correction, the no-pair Coulomb all-order dipole matrix elements for resonant transitions are brought into an excellent agreement with the accurate experimental values. This work was supported by the U.S. Department of Energy, Division of Chemical Sciences, Office of Energy Research. The author would like to thank W.R. Johnson for useful discussions and H.R. Sadeghpour for suggestions on manuscript.
warning/0001/astro-ph0001286.html
ar5iv
text
# Starbursts in barred spiral galaxies ## 1 Introduction Massive galaxies play a fundamental role in the chemical evolution of the Universe, because they are the only systems which meet the physical conditions necessary for the build–up of heavy elements. Today, by using simple techniques and modest–size telescopes, it is possible to understand how the production of metals occurs in nearby massive galaxies and to relate this production to their formation and evolution processes. While still in its early stages, the study of chemical abundances in external galaxies has already produced important results. One of the most interesting is the discovery that galaxies in the Universe seem to follow a mass–metallicity relation Zaritsky et al. (1994). Although we still have to understand the physical causes behind this behavior, there is no doubt that this phenomenon is a key parameter for understanding galaxy evolution. With this prospect in mind, we recently verified that the massive starburst nucleus galaxies (SBNGs) follow the same mass–metallicity relation as normal ones Coziol et al. (1997a). Our observations not only confirm the universality of the mass–metallicity relation for a wide range of galaxies, but also suggest that its origin may be linked to the different paths followed by early and late–type spiral galaxies to form their bulge and disk Coziol et al. (1998). Another important phenomenon, unveiled by chemical abundance studies, is that the metallicity in the ionized interstellar medium of spiral galaxies decreases outward (e.g. Vila–Costas & Edmunds, 1992, and references therein). Various hypotheses have been proposed to explain this gradient Pagel (1989); Götz & Köppen (1992), but their verification is difficult. The reason is that many parameters and processes involved in the formation of an abundance gradient are still not well understood, such as the initial mass function, the yield, the mechanisms and time scales for the formation of halos, bulges and disks, as well as the possible infall or outfall of matter. Different numerical simulations have shown, however, that one important ingredient for producing abundance gradients is a non linear dependence of the star formation rate on gas density Wise & Silk (1989); Götz & Köppen (1992); Mollá et al. (1996). It has also been suggested that once an abundance gradient is established, the presence of gas flows can either amplify or reduce it Edmunds (1990); Götz & Köppen (1992). Attempts to relate abundance gradients to other characteristics of galaxies, such as morphological type, luminosity or circular velocity, have also been performed Vila–Costas & Edmunds (1992); Oye & Kennicutt (1993); Zaritsky et al. (1994), but no clear correlation has emerged. One interesting trend is however observed: barred spiral galaxies seem to have shallower abundance gradients than non–barred ones Vila–Costas & Edmunds (1992); Edmunds & Roy (1993); Zaritsky et al. (1994); Martin & Roy (1994). From a theoretical point of view, bars are expected to funnel gas from the outer parts to the nuclei of galaxies Noguchi (1988). The flow along the bar can induce strong mixing Friedli et al. (1994) and reduce metallicity gradients. But star formation induced by the bar can also be strong enough to counteract dilution effects and increase the gradient in the inner parts of galaxies. Steep abundance gradients have indeed been reported in the inner parts of some barred galaxies, such as NGC 3359 Martin & Roy (1995) and NGC 1365 Roy & Walsh (1998). In the present paper, we analyse the abundance gradients of oxygen (O/H) and nitrogen to oxygen ratio (N/O) in 16 barred starburst galaxies. Our goal is to establish if starburst galaxies follow the trends observed in normal galaxies. We also want to determine if there is a relation between the bar and the burst of star formation. We have recently proposed that SBNGs may be “young” galaxies still in their formation process (Coziol et al. 1997a,1998). This possibility raises several new questions. Is the initial distribution of abundances in starburst galaxies flat, or have they already had enough time to establish steep gradients? Is it possible to determine how the gradient builds up in a young galaxy and establish whether the presence of the bar is important? Could bars be younger in starburst than in normal galaxies? Our analysis innovates by studying the abundance gradients of two key elements, oxygen and nitrogen, which are assumed Coziol et al. (1999a) to be produced by different types of stars and, thus, to be released into the interstellar medium on different time scales. This gives us a deeper insight into the processes of chemical evolution, because we can test the influence of star formation and of structures, like a bar, over a relatively long period of time (a few Gyrs). This advantage should in particular enable us to determine whether bars are young and if they play a major role in triggering or feeding the nuclear starbursts. ## 2 The sample of barred starburst galaxies The 16 galaxies studied in this paper are a sub–sample of 144 Markarian barred spiral galaxies studied spectroscopically by Contini Contini (1996) and Contini et al. Contini et al. (1998) (Paper III). This sub–sample is a selection of galaxies which are experiencing very intense bursts of star formation, as estimated by their high H$`\alpha `$ luminosity (L(H$`\alpha `$) $`>10^{40}`$ erg s<sup>-1</sup>) or large H$`\beta `$ equivalent width (EW(H$`\beta `$) $`>`$ 30 Å). Three Wolf–Rayet galaxies are included in our sample: Mrk 710, 712 and 799. Another Wolf–Rayet galaxy satisfying our selection criteria, NGC 6764, Contini et al. (1997), was later added to the list; but, as we will show later, it turns out to be a LINER and cannot be included in our analysis. Some catalog elements for the sample galaxies and excerpts from our observing logbook are presented in Table 1. The morphologies were taken in NED (http://nedwww.ipac.caltech.edu), or in LEDA (http://www–obs.univ–lyon1.fr) for Mrk 13 and 306. Table 1 also lists the distance $`D`$, estimated from the recession velocity (assuming H$`{}_{o}{}^{}=75`$ km s<sup>-1</sup> Mpc<sup>-1</sup>), and the radius $`r_{25}`$ at 25 mag arcsec<sup>-2</sup>. Both values were found in LEDA. The last two columns indicate the exposure times and position angles of the slit which were used during our spectroscopic observations. ## 3 Spectroscopic observations and reductions Most of the spectroscopic observations were obtained in October 1995 and June 1996 at the 1.93 meter telescope of Observatoire de Haute–Provence. The CCD was a thinned 512$`\times `$512 Tektronix (pixel size 27$`\mu `$m). We used the Carelec spectrograph Lemaître et al. (1990) with a spectral dispersion of 260 Å/mm, which covers the spectral range 3600 to 7200Å, at a resolution of $`7`$Å. The slit was aligned along the bar. During the first run, we took 90–minute spectra of most galaxies, with a slit width of 3.0″. For flux calibration, we observed the standard stars G191B2B, Hilt 600, HD 217086 and BD +28 4211 Massey et al. (1988). In the second run, we took three 25–minute spectra of Mrk 799 and two of NGC 6764, with a slit width of 2.8″. One standard star, HD 192281, was used for flux calibration Massey et al. (1988). The galaxies Mrk 710 and 712 were observed during previous runs at the same telescope, using similar instrumental settings Contini et al. (1998). The average seeing during the nights was of the order of 1.2″. All spectra were reduced with MIDAS, applying standard procedures: offset and flatfield corrections, sky subtraction, wavelength and flux calibration, airmass and galactic extinction corrections and elimination of cosmic impacts. In order to study abundance variations along the bar, series of one–dimensional spectra were extracted from the two–dimensional spectra of each galaxy by averaging three contiguous rows along the slit. Observed emission–line fluxes and equivalent widths were then measured in each elementary spectrum. For the rest of the reduction we followed the procedure outlined in Contini et al. Contini et al. (1995) (Paper I): the Balmer (H$`\alpha `$ and H$`\beta `$) absorption contamination due to underlying stellar populations was corrected by adding 2Å to the equivalent width of the Balmer emission lines; the principal Balmer decrement (H$`\alpha `$/H$`\beta `$) was used to estimate the reddening coefficient $`c_{\mathrm{H}\beta }`$ and to calculate absolute dereddened fluxes. Our analysis may differ from other studies found in the literature, because we measured the spatial variations of physical parameters only along one direction (along the bar), and not across the whole two–dimensional disk. This difference should be kept in mind when comparing our results with those of other studies. ## 4 Properties of emission–line regions A standard diagnostic diagram Baldwin et al. (1981); Veilleux & Osterbrock (1987) was used to determine the dominant source of excitation of the emission–line regions in our sample. Most of these regions fall in the domain of Hii region–like spectra (Fig. 1), referred to hereafter as Hii regions. The only exceptions are two regions in Mrk 545, which have weak signal–to–noise ratios, and the inner region of NGC 6764. There has been some confusion in the literature about the nature of the activity in this last galaxy (see Contini et al. 1996 for the various classifications). Our new spectra clearly show that the center of NGC 6764, as well as the circum–nuclear regions, are the seat of LINER activity. Because LINER is a non–thermal phenomenon (implying either an AGN or shocks), the metallicity in these regions cannot be determined by the methods used in this paper. Consequently, NGC 6764 (and the two regions in Mrk 545) have been excluded from our abundance gradient analysis. Our sample of Hii regions spans a large range of excitation levels, from high–excitation spectra (log (\[O iii\]/H$`\beta `$) $``$ 0.4), characteristic of Hii galaxies, to low–excitation spectra (log (\[O iii\]/H$`\beta `$) $`<`$ 0.4) typical of SBNGs Coziol (1996). The low–excitation Hii regions have \[N ii\]$`\lambda `$6584/H$`\alpha `$ ratios which are relatively high. This is a common property of SBNGs Coziol et al. (1997b); Contini et al. (1998), which possess a slight overabundance of nitrogen as compared to Hii regions with comparable metallicity Coziol et al. (1999a). ## 5 Determination of oxygen and nitrogen abundances ### 5.1 Determination of oxygen abundances: general method The oxygen abundances are determined using empirical relations between log(O/H) and the emission–line ratios $`R_3`$ and $`R_{23}`$, which are defined as: $`R_3=(\text{[O }\text{iii}\text{]}\lambda \text{4959}+\text{[O }\text{iii}\text{]}\lambda \text{5007})/\text{H}\beta `$ (1) $`R_{23}=(\text{[O }\text{ii}\text{]}\lambda \text{3727}+\text{[O }\text{iii}\text{]}\lambda \text{4959}+\text{[O }\text{iii}\text{]}\lambda \text{5007})/\text{H}\beta `$ (2) For our analysis we use the empirical relations first established by Edmunds & Pagel Edmunds & Pagel (1984) and quantified by Vacca & Conti Vacca & Conti (1992). In general, the above two relations yield similar results. In some cases, however, the relation based on $`R_3`$ gives higher oxygen abundances than the one based on $`R_{23}`$, because the extinction correction affects differently the two relations; a high extinction correction tends to decrease the ratio $`R_3`$, while increasing the ratio $`R_{23}`$. The consequence of this effect is that when dust extinction is high, one overestimates the oxygen abundance using $`R_3`$ and underestimates it using $`R_{23}`$. For this reason, we will use both relations and adopt the mean value. Exceptions are Mrk 710 and 712, where \[O ii\]$`\lambda `$3727 was not observed and only $`R_3`$ was used. ### 5.2 Determination of extinction in the center of galaxies In the center of half of the sample galaxies (Mrk 12, 307, 326, 332, 545, 799, 898 and 1088), Balmer absorption features due to an underlying intermediate–age stellar population severely affect the measurement of the Balmer emission lines. This prevents us from determining the extinction coefficient $`c_{\mathrm{H}\beta }`$ which is needed for calculating the abundances. The standard correction of 2Å on the equivalent width is not adequate for these cases (because we still cannot measure H$`\beta `$ correctly) and it is thus necessary to find a more suitable method to solve this problem. It has been known for some time that there is an empirical relation between log$`(R_3)`$ and log(\[N ii\]/\[O iii\]) Alloin et al. (1979); Edmunds & Pagel (1984). The advantage of using the \[N ii\] and \[O iii\] emission lines for determining the oxygen abundance is that they are not affected by underlying stellar populations. Such a method is ideal for analyzing the central regions of galaxies. However, we still have to correct these lines for dust extinction. As we will now show, this can be done empirically. According to the definition of the extinction coefficient, the observed and dereddened fluxes of the \[N ii\]$`\lambda `$6584 and \[O iii\]$`\lambda `$5007 lines (hereafter \[N ii\] and \[O iii\]) are linked by the relation: $`\mathrm{log}\text{[N }\text{ii}\text{]/[O }\text{iii}\text{]}_{dred}=\mathrm{log}\text{[N }\text{ii}\text{]/[O }\text{iii}\text{]}_{obs}0.3\times c_{\mathrm{H}\beta }`$ (3) If one plots log(\[N ii\]/\[O iii\])<sub>dred</sub> versus log(\[N ii\]/\[O iii\])<sub>obs</sub> for a sample of spectra, statistically large enough to include evenly distributed values of $`c_{\mathrm{H}\beta }`$ between 0.0 and $``$ 1.5, one expects the data to cover uniformly the region between the line of zero extinction and that of $`c_{\mathrm{H}\beta }`$ $``$ 1.5. We notice instead that the data gather along an oblique line of slope smaller than unity, which intersects the lines of constant $`c_{\mathrm{H}\beta }`$ (see Fig. 2). This is observed in two samples of Hii regions in normal galaxies McCall et al. (1985); Vila–Costas & Edmunds (1993), as well as in our own sample of starburst galaxies (excluding the spectra affected by Balmer absorption). This correlation translates the fact that, on average, the larger the value of log(\[N ii\]/\[O iii\]) the larger that of $`c_{\mathrm{H}\beta }`$. The reason behind this behavior is that log(\[N ii\]/\[O iii\]) is correlated with oxygen abundance (see Sect.5.3 and Fig. 3), which in turn is correlated with the amount of dust and consequently with internal extinction Heckman et al. (1998). This observed correlation provides us with an empirical way of estimating $`c_{\mathrm{H}\beta }`$ from the observed fluxes of \[N ii\] and \[O iii\]. Using the sample of galaxies of McCall et al. (1985) and our own sample (excluding the spectra with Balmer absorption), we find the following relation between the observed and dereddened line ratios: $`\text{log([N }\text{ii}\text{]/[O }\text{iii}\text{])}_{dred}=0.93\text{log([N }\text{ii}\text{]/[O }\text{iii}\text{])}_{obs}0.16`$ (4) The correlation coefficient of this relation is 0.99 and the uncertainty on log(\[N ii\]/\[O iii\])<sub>dred</sub> is less than 0.1 dex. This relation will be used in Sect. 5.3 to calculate (\[N ii\]/\[O iii\])<sub>dred</sub>, and thus $`R_3`$, in the center of galaxies affected by Balmer absorption. Using the same samples, we find the following relation between $`c_{\mathrm{H}\beta }`$ and (\[N ii\]/\[O iii\])<sub>obs</sub>: $`c_{\mathrm{H}\beta }=0.45\text{log([N }\text{ii}\text{]/[O }\text{iii}\text{])}_{obs}+0.51`$ (5) The correlation coefficient is about 0.95 and the uncertainty is about 0.1 in $`c_{\mathrm{H}\beta }`$. This relation will be used in Sect. 5.4 to deredden the \[O iii\]$`\lambda `$5007, \[O ii\]$`\lambda `$3727 and H$`\alpha `$ lines in the center of galaxies with strong Balmer absorption, and hence to calculate $`R_{23}`$. ### 5.3 Calculating $`R_3`$ in the center of galaxies Following an idea developed by various authors Alloin et al. (1979); Edmunds & Pagel (1984), we now search for an empirical relation between $`R_3`$ and \[N ii\]/\[O iii\] in our sample of 16 barred starburst galaxies. For this analysis we use only the dereddened fluxes of Hii regions, about 300 of them, which are not affected by Balmer absorption. The two quantities appear to be strongly correlated, as shown on Fig. 3a, and we fit a two–degree polynomial to the data, which yields the following relation: log$`(R_3)`$ $`=`$ $`0.123(\text{log([N }\text{ii}\text{]/[O }\text{iii}\text{])})^2`$ (6) $`0.661\text{log([N }\text{ii}\text{]/[O }\text{iii}\text{])}+0.086`$ with a dispersion in log$`(R_3)`$ equal to 0.07 dex. To check the validity and the possible “universality” of the above relation, we now apply the same procedure to two samples of Hii regions in normal galaxies: those of McCall et al. (1985) and of Vila–Costas & Edmunds (1993). As shown in Fig. 3b, the behavior is the same. Fitting a two–degree polynomial we obtain: log$`(R_3)`$ $`=`$ $`0.133(\text{log([N }\text{ii}\text{]/[O }\text{iii}\text{])})^2`$ (7) $`0.709\text{log([N }\text{ii}\text{]/[O }\text{iii}\text{])}0.008`$ The dispersion in log$`(R_3)`$ is 0.12 dex, that is slightly higher than for our sample. It is important to note that the two fitted curves in Fig. 3 differ by a horizontal shift of 0.14 dex. This shift is consistent with the overabundance of nitrogen in SBNGs with respect to normal Hii regions Coziol et al. (1999a). Consequently, there is no unique relation between log$`(R_3)`$ and log(\[N ii\]/\[O iii\]), but specific relations for homogeneous samples of galaxies, such as normal ones or SBNGs. For a given value of the ratio \[N ii\]/\[O iii\], a starburst galaxy will have a lower oxygen abundance than a normal galaxy, or, for a given abundance, a starburst galaxy will have a higher \[N ii\]/\[O iii\] ratio than a normal galaxy. Figure 3a shows that our data do not reach oxygen abundances as high as in normal galaxies. This is precisely the domain where we cannot determine $`R_3`$ in the standard way, because, in starburst galaxies, such high values for oxygen abundance are observed only in the nuclear regions, which are affected by Balmer absorption features. Fortunately, the parallel empirical relation for normal Hii regions (Fig. 3b) continues smoothly into the high oxygen–abundance regime. We therefore consider it legitimates to use Eq.6 to calculate $`R_3`$ in this range of oxygen abundances. ### 5.4 Calculating $`R_{23}`$ and O/H in the center of galaxies Once we know $`c_{\mathrm{H}\beta }`$ in the center of the sample galaxies (using Eq. 5), it is relatively easy to estimate $`R_{23}`$. First, we correct \[O iii\]$`\lambda `$5007, \[O ii\]$`\lambda `$3727 and H$`\alpha `$ for reddening using $`c_{\mathrm{H}\beta }`$ and estimate the dereddened H$`\beta `$ flux using the theoretical Balmer decrement: H$`\beta `$$`{}_{dred}{}^{}=`$H$`\alpha `$$`{}_{dred}{}^{}/2.85`$. Then, the dereddened ratio $`R_{23}`$ is computed assuming \[O iii$`=1.35\times `$ \[O iii\]$`\lambda `$5007, and the oxygen abundance O/H is deduced using the relation of Vacca & Conti Vacca & Conti (1992). Comparing the abundances obtained using $`R_3`$ and $`R_{23}`$ in the center of galaxies, we find the same kind of sensitivity to dust extinction as that observed in the outer parts of galaxies. Consequently, we also adopt the average value of the oxygen abundances estimated by $`R_3`$ and $`R_{23}`$ in the center of our sample galaxies. ### 5.5 N/O abundance ratios With the data at hand, it is possible to derive the nitrogen–to–oxygen abundance ratio (N/O). For Hii regions in our sample, the N/O abundance ratio has been derived from the dereddened emission lines \[O ii\]$`\lambda `$3727, \[O iii\]$`\lambda `$5007 and \[N ii\]$`\lambda `$6584 following the method of Thurston et al.Thurston et al. (1996), as described in Coziol et al. Coziol et al. (1999a). The N/O abundance ratio as a function of oxygen abundance is shown in Fig. 4. Also shown in this figure are the expected relations for a $`primary`$, a $`secondary`$ and a $`primary+secondary`$ origin for the production of nitrogen Vila–Costas & Edmunds (1993). The Hii regions in our sample seem to follow the $`primary+secondary`$ relation. But the evolution of N/O with oxygen abundance does not trace a continuous behavior: it suddenly rises and forms a sort of plateau between $`8.7<12+\text{log(O/H)}<9.1`$. This behavior has been interpreted as evidence for chemical evolution by a sequence of bursts of star formation (Coziol et al. 1999a). ## 6 Results ### 6.1 Spatial distribution of oxygen abundances and of H$`\alpha `$ The oxygen abundance and H$`\alpha `$ emission as a function of position along the slit are presented in Fig. 5. The emission peak is usually centered on the nucleus of galaxies. Exceptions are Mrk 12, 13, 712 and 799. In the case of Mrk 712 and 799, the maximum intensities correspond to very young star–forming regions containing numerous Wolf–Rayet stars. When studying the gas distribution inside the bar, we find two possible situations: more than half of galaxies have ionized gas covering the whole bar, while in the rest of galaxies the ionized gas is mostly confined to the center. A large number of galaxies (62%) also show ionized gas either at the two ends of the bar (25%), or, more frequently, only at one end (37%). Examining the spatial distribution of the oxygen abundance, we find that it usually reaches a maximum in the center of galaxies. In general, this maximum corresponds to that of the H$`\alpha `$ emission. In some galaxies, there is also strong nebular emission outside the nucleus, accompanied by a local rise in oxygen abundance (e.g. Mrk 306, 332 and 545). In other galaxies this secondary peak of emission is not accompanied by an increase in oxygen abundance (e.g. Mrk 12, 13 and 799), and in some galaxies the oxygen abundance stays high even without any secondary peak of H$`\alpha `$ emission (e.g. Mrk 307, 373 and 712). ### 6.2 O/H and N/O abundance gradients To study the abundance gradients, we have folded the profiles of O/H and N/O abundance ratios about a center of symmetry. This center is the locus of the principal peak of H$`\alpha `$ emission (see Fig. 5). To make the analysis easier, we adjusted lines by least–square fits to the data in two different regions of each profile: one includes the bar and the nucleus and another the region outside the bar. The latter region corresponds to the disk. Note, however, that the profiles never cover the entire length of the disk (that is out to a surface brightness of 25 mag arcsec<sup>-2</sup>), simply because we did not detect ionized gas that far. A global gradient is also estimated by fitting a line to the data over the full extent of the ionized gas distribution. The intersect (central) abundance is determined by extrapolating this global gradient up to the center. The two regressions are plotted as continuous lines and the global gradient as a dotted line in Figs. 6 and 7. The profiles on the two sides of the center are distinguished by different symbols. The numerical values of different gradients and the intersects are given in Table 2. The uncertainties reported in that Table are those obtained from the regression calculations. The oxygen abundance profiles as a function of radius (in arcsec) are shown in Fig. 6. They show complicated patterns. Although the uncertainties on the oxygen abundance ($`0.2`$ dex) prevent us from interpreting the multiple inflections of these profiles, we believe that part of these features must be real, reflecting variations in oxygen abundance produced by different intensities of star formation along the bar. One good example is Mrk 332, where there is an increase of oxygen abundance linked to star formation at the two ends of the bar. The abundance gradients (continuous and dotted lines in Fig. 6) also show various patterns. In Mrk 2, 373, 710, 712, 898 and 1076, the gradient can be estimated in the bar only. In these cases, the bar gradient is considered the global gradient. Half of these gradients are relatively shallow ($`0.03`$) and half are moderately strong ($`0.15`$). For galaxies with ionized gas also in the disk, we distinguish two cases: the disk abundance gradient can be either flat or negative. In Mrk 306, 326, 545 and 602, it is flat. Considering that normal spiral galaxies usually have monotonous negative gradients Zaritsky et al. (1994), this makes these galaxies with flat disk gradient somewhat peculiar, although shallower outer profiles have been observed before, for instances in NGC 1365 Roy & Walsh (1998), NGC 3319 Zaritsky et al. (1994) and NGC 3359 Martin & Roy (1995). Other galaxies with ionized gas in the disk look more “normal”, with negative disk gradients. In general, the gradients are shallower in the disk than in the bar. There are two exceptions, Mrk 13 and Mrk 332. In the case of Mrk 332, the intense star formation located at the ends of the bar is obviously responsible for the steeper gradient in the disk (see Fig. 6). For Mrk 13, there is no gradient in the bar simply because this bar is very short, it is the shortest bar among the 125 ones measured by Chapelon et al. Chapelon et al. (1999) (paper V). The bar of Mrk 1088 also seems to have a flat gradient (Fig. 6), but the unfolded oxygen abundance profile (Fig. 5) shows that there is in fact a steep gradient accross the center. In general, our global abundance gradients are negative. This reflects the fact that the oxygen abundance is almost always highest in the center of the galaxy. One exception is Mrk 712, which does not have any gradient at all. We determined the N/O abundance ratios of all galaxies except Mrk 710 and 712 (see Sect. 5.5). The N/O profiles are shown in Fig. 7. In general, they look similar to those of oxygen. The N/O abundance ratio also increases toward the nucleus of the starburst galaxies, but the gradients are significantly shallower and many more galaxies have zero gradient. Further comparisons of the two abundance gradients are postponed to Sect. 7. Values for the different gradients and the intersects are given in Table 2. ## 7 Analysis of the abundance gradients ### 7.1 Oxygen abundance gradients The presence of a bar could affect the abundance gradients in our sample of galaxies in two ways. By funneling metal–poor gas from the outer regions toward the center, the bar is expected to dilute the chemical elements, and thus reduce any initially present radial abundance gradient. But a bar could also stimulate star formation at the centre, which, if strong enough, could establish or maintain a steep abundance gradient. In the following analysis, we look for such possible effects of the bar in our sample of starburst galaxies. Of the three categories of oxygen abundance gradients estimated here, the global gradient is the one which is most comparable to those from other studies. We show in Fig. 8 the relation of global gradients and intersects with absolute magnitudes. This Figure should be compared with Fig.1 of Edmunds & Roy (1993) for a sample of normal galaxies. The barred starburst galaxies generally have steeper gradients than normal barred galaxies Edmunds & Roy (1993); Zaritsky et al. (1994). On average, the gradients in the barred starburst galaxies are even steeper than those measured by Edmunds & Roy (1993) in normal unbarred galaxies of comparable luminosity. This shows that the oxygen abundance in the sample of barred starburst galaxies is not diluted by the mixing effect of the bar. In the above analysis, six of the global abundance gradients are in fact bar gradients, because no ionized gas was observed in the disks of these galaxies. We should therefore reexamine Fig. 8 and distinguish these cases from those where the global gradient includes the disk gradient. With this distinction, we find that the global gradient in Fig. 8a tends to steepen as the absolute magnitude increases. A linear fit performed on the data yields a correlation coefficient of 86%. But if we exclude the extreme point (Mrk 13) at $`M_\mathrm{B}18`$, this coefficient drops to 56%. This is more a trend than a true correlation. Galaxies for which the global gradient is the bar gradient also show a similar trend, but slightly displaced towards lower values. An explanation for this trend is proposed in Sect. 8.2. Examining now the central oxygen abundances (Fig. 8b), we find that they are on average lower in barred starburst galaxies than in normal unbarred galaxies. Because the global abundance gradients in barred starburst galaxies are strong, these low central abundances cannot be attributed to an artifact of the bar, that would have lowered the central abundance of chemically evolved galaxies by dilution. This means that the starburst galaxies in this sample have lower oxygen abundances than normal ones. This is in agreement with our previous finding Coziol et al. (1997a, 1998): SBNGs are chemically less evolved than normal galaxies with comparable luminosity and morphology. We have shown elsewhere Coziol et al. (1997a, 1998), that the only way to explain this phenomenon is by assuming that SBNGs are “young” galaxies. ### 7.2 N/O abundance gradients We have shown that trends in the oxygen abundance gradients are inconsistent with the mixing effect of a bar, unless the bar also triggers a central starburst which more than compensates for the dilution. However, the fact that we find both strong oxygen abundance gradients and lower intersects than in normal galaxies rules out a bar–triggered burst in the center of an “old” galaxy. But the bar could be responsible for starbursts in the center of a relatively “young” galaxy. In this case, there are two alternatives. The gas funneled by the bar could have been fed directly into the central region in a short time (a few Myrs), prompting a new burst of star formation that is sufficiently strong to produce steep abundance gradients. The other alternative is that the bar acts over a much longer time scale (a few Gyrs), progressively feeding the inner regions with gas that is immediately recycled (since no mixing effects are seen). We can test the above two alternatives by studying the N/O abundance gradients and comparing them with the oxygen ones. In Coziol et al.Coziol et al. (1999a), we have shown that the bulk of nitrogen in SBNGs is probably produced by intermediate–mass stars, whereas oxygen is mainly released by massive stars. In this scenario, the time scale for enrichment in nitrogen ($`400\text{Myr}<\tau <2\text{Gyr}`$) is much longer than the one for oxygen ($`\tau <20`$ Myr). This allows us to look for effects of the bar on a time scale covering a few Gyrs. In Fig 9a, we show the relation between the global N/O abundance gradient and absolute magnitude. In this Figure, we distinguish galaxies with a gradient measured only along the bar from those where the gradient also includes the disk. We observe the same trend for N/O as for oxygen: galaxies with a lower luminosity seem to have a stronger negative gradient. The relation implied by this tendency has a weak correlation coefficient (60% with Mrk 13 excluded) and a shallower slope than that found for oxygen. Again, galaxies with only a bar gradient seem to have shallower gradients than the other ones. Despite the small gradients, the central N/O values, shown in Fig. 9b, are rather high compared to the solar abundance. This implies that many generations of intermediate–mass stars have already contributed to enrich these galaxies in nitrogen, or, equivalently, that enough time has passed in the chemical evolution of these galaxies. Why is the gradient systematically weaker for N/O than for oxygen? If this is due to the bar, it means that the effect of the bar appears on a time scale comparable to the nitrogen enrichment phase. This is unlikely as it implies that the oxygen enrichment phase, and the ensuing strong oxygen abundance gradient, happened after the nitrogen one. Therefore, as a mixing mechanism, the bar should have affected both elements in the same way, which is not observed. Our explanation for the observed difference is given in Sect 7.3. The fact that the N/O intersects are high rules out the possibility that the bars recently triggered central starbursts in these galaxies. A recent burst would have increased the oxygen abundance and decreased the N/O abundance ratio. The stronger the burst, the stronger the decrease in N/O. We would thus expect galaxies with the steepest oxygen abundance gradients to have the flattest N/O ones. In Fig. 10a, we show the relation between the oxygen and N/O abundance gradients. The behavior is contrary to what is expected. The bars consequently did not recently trigger bursts in these galaxies. ### 7.3 The progressive build up of abundance gradients What is then the reason for the difference between the two gradients? In fact, the N/O abundance gradient is not independent of the oxygen one. This is shown in Fig. 10a. A linear regression on the data yields the relation: $`\mathrm{\Delta }(\mathrm{N}/\mathrm{O})/\mathrm{\Delta }\mathrm{r}=0.53(\pm 0.10)\mathrm{\Delta }(\mathrm{O}/\mathrm{H})/\mathrm{\Delta }\mathrm{r}+0.00(\pm 0.01)`$ with a correlation coefficient of 89%. Surprisingly, the slope of this relation is almost the same as the one found between the N/O and oxygen abundances: a linear fit performed on the data in Fig. 4 yields log(N/O)$`=0.55`$log(O/H)$`+0.8`$ Coziol et al. (1999a), which is consistent with a mixture of $`primary+secondary`$ origin for nitrogen McGaugh (1991). The relation between the two gradients is thus explained by the behavior of the oxygen abundance in the nuclei of the galaxies. This is shown in Fig. 10b, where the oxygen abundance gradient steepens when its intersect increases. This relation is trivial if we assume that all disks initially have about the same low oxygen abundance (see Fig. 6). Since the nitrogen enrichment is related to the increase in oxygen (See Sect. 5.5 and Fig. 4), the N/O abundance gradient grows with that of oxygen. The origin of both gradients in starburst galaxies can thus be explained solely by their star formation histories. ## 8 The role of bars in starburst galaxies The only way to reconcile the above observations of abundance gradients with the possible effect of a bar is by assuming that the latter is a slow process; over a few Gyr period, the bar progressively feeds the nucleus with gas which is immediately recycled into stars. This scenario requires a very fine tuning between the feeding of the gas in the center of galaxies and star formation. It also requires that bars are relatively stable over a long period of time (a few Gyrs). We can test if these conditions apply to galaxies in our sample by looking for relations between the bar properties and star formation. ### 8.1 Bar properties and star formation: evidence for young bars Numerical simulations predict that a strong bar maintains a steeper abundance gradient in the bar than in the disk for about 1 Gyr, then both become comparably low ($`0.02`$), and that the bar and disk gradients are never very different in the presence of a weak bar Martinet & Friedli (1997). We show in Fig. 11 that almost all galaxies in our sample show a stronger gradient in the bar than in the disk. Only two galaxies, Mrk 13 and 332, have significantly steeper gradients in the disk than in the bar, but this peculiarity can be explained by other phenomena (see Sect. 6.2). In other words, the difference between bar and disk gradients cannot be due to an old bar. It cannot be due to the effect of a young bar either, because, as we previously noted, the N/O abundance ratios would be low and the gradients would be anti–correlated. A simple explanation for the stronger gradients in the bar than in the disk is that the volume density of star formation is proportional to the gas volume density to some power $`n`$ ($`1n2`$), as proposed by Schmidt’s law (1959, 1963). Assuming that the chemical enrichment is proportional to the star formation, a spherical density distribution yields a variation with radius in $`R^{3/n}`$. The gradient measured from the center to the end of the bar (or to any intermediate radius) will necessarily be steeper than the one measured further out. The presence of ionized gas is a tracer of star formation in galaxies. Half the galaxies in this sample show the presence of ionized gas over the whole bar, while in the other half, the gas is mostly concentrated in the inner parts of the bar and in the nucleus. If star formation cannot result from gas flow toward the inner regions by the dynamical action of a bar, could it be linked to any other property of the bar, such as its strength or its length? The bar strength is identified with the deprojected bar axis ratio (b/a)<sub>i</sub>, and its length is given by the ratio of the deprojected major axis to corrected blue isophotal diameter 2a/D<sub>c</sub> Chapelon et al. (1999) (Paper V). A bar with a ratio b/a $`<0.5`$ is considered “strong” and a bar with a ratio 2a/D$`{}_{c}{}^{}>0.2`$ is considered “long”. We also identify separately galaxies with various fractions of ionized gas along the bar and galaxies with early– (earlier than SBbc) or late–type (SBbc or later) morphologies. As shown in Fig. 12, most galaxies in this sample have a strong bar, which is a general characteristic of the sample of Markarian barred starburst galaxies Chapelon et al. (1999). There is a trend for galaxies with a longer bar to have ionized gas over a smaller extent of the bar length, regardless of their morphology. This is not an observational selection effect, because three galaxies do not follow the above trend in Fig. 12, namely Mrk 306, 712 and 1076. In the case of Mrk 306 and 1076, a long and strong bar with ionized gas over its whole length may be explained by the fact that they are both interacting with close companions (the rest of our galaxies are isolated). Mrk 712 looks like an isolated galaxy, but with a peculiar morphology. The peak of star forming activity is significantly displaced from the center of the galaxy (see Fig. 5). This active region may be associated with a second nucleus Mazzarella & Boroson (1993), which makes this galaxy a strong candidate for a recent merger. A simple relation with the length of the bar thus explains the radial distribution of ionized gas; star formation does not cover the whole bar just because it is too long. We have shown so far that bars, despite the fact that they are strong, cannot account for the observed abundance gradients. This may be surprising in view of the convincing evidence for dynamical effects of strong bars on star formation and abundance gradients in galaxies provided by numerical simulations. The only explanation for such a situation is that bars in the current sample of galaxies are too young (a few $`10^7`$ years) to have had any effect. We verified that there are no relations between the abundance gradients and the bar properties. The oxygen and N/O abundance gradients in the bar and in the disk show the same behavior. No relation is found between the abundance gradients and the bar strength or length. Nor is any relation found between the abundance gradients and the different concentrations of gas along the bar. In general, therefore, the bars have not influenced the two gradients. This supports our assumption that the bars are young. ### 8.2 Bar–induced star formation at the bar ends If we compare the distribution of ionized gas (i.e. star formation) with the oxygen abundance, we find that the behaviors of these two parameters along the bars are frequently uncorrelated (see Fig. 5 and Sect. 6.1). Intense star formation does not necessarily enhance the oxygen abundance, as we can see, for example, in the extranuclear star forming regions of Mrk 12 and 799. These regions must be too young (a few Myrs) and their massive stars did not had enough time to change the chemical abundance of their environment. When we find, on the other hand, a region filled with ionized gas, but where the abundance is significantly higher, we conclude that this must be a region where star formation was “stable” over a longer period of time. The galaxy nuclei in our sample are obvious locations for such stable star-forming regions. The relatively high N/O abundance ratios and the high proportion of intermediate–mass stars mixed with ionized gas indicate that star formation persisted in these regions for quite a long time (a few Gyrs). We found no evidence that bars triggered the starbursts observed at the center of the sample galaxies. On the other hand, they may have induced star formation at the ends of the bar (e.g. Mrk 307, 332), and more frequently at only one end (e.g. Mrk 12, 13, 306, 545 and 799). This phenomenon may be due to gas compression in these regions Roberts et al. (1979). Star formation induced by the bar at its ends has increased the oxygen abundance there. The result is a shallower abundance gradient in the bar and a stronger one in the disk. This may explain why the global oxygen and N/O abundance gradients are weaker when they are measured only in the bar. Chemical enrichment produced by star formation at these points being lower than what is observed in the nucleus, this suggests that fewer star formation episodes occurred there, or that they happened relatively recently. Among all the possible effects of the bar on the evolution of starburst galaxies, therefore, the only plausible one seems to be star formation induced at one or both ends of the bars. The distribution of star formation in this sample of starburst galaxies is consistent with the assumption that bars are too young (a few $`10^7`$ years old) to be at the origin of the bursts in the galaxy nuclei. ## 9 Summary and conclusion One important result of our analysis is that bars have not had a great impact on the star formation history and chemical evolution of starburst galaxies. The most straightforward explanation is that bars are too young. Our observations also clearly show that these young bars cannot be at the origin of the nuclear starbursts. The high N/O abundance ratios together with the high level of star formation in the nuclear region of galaxies suggest that star formation proceeds almost steadily (or as a sequence of bursts) over a few Gyr period. The fact that we find strong O/H gradients while the oxygen abundance is low in the galactic nuclei supports our interpretation that SBNGs are young galaxies still in their process of formation (Coziol et al. 1997a; 1997b). It seems that these starburst galaxies are still building their abundance gradients. This process can be fully explained in terms of their star formation history. The gradients build up from the inside out, becoming stronger as the oxygen and N/O abundances increase in the bulge while staying low in the disk. This behavior is consistent with a simple Schmidt law relating the density of star formation to that of gas. According to these results, starburst galaxies seem to form their bulge first, and then their disk. If the disk is younger than the bulge and the bar forms in the disk, it is not surprizing that we find mostly young bars in starburst galaxies. The above scenario may also explain why our results are not consistent with predictions made by different models of bar formation. The initial conditions assumed in these models are far from those observed in young starburst galaxies. A bar forms in a galaxy which already has a large bulge but a disk which is probably still in formation. When star formation stops in the bulge and increases in the disk, as in normal spiral galaxies, the metallicity of the disk grows, decreasing the gradient. If the end product of a starburst galaxy is a normal spiral galaxy, the latter, therefore, should have lower metallicity gradients than the former. If they are stable, bars may also play a more important role as galaxies get older. But how general is our conclusion? Is what we observe a trait of SBNGs? What is the relation between the bulge and disk formation in starburst galaxies and in normal spiral galaxies? These are the questions we address in our companion paper (Coziol et al. 2000). ###### Acknowledgements. We thank the staff of Observatoire de Haute–Provence for assistance at the telescope. R. C. would also like to thank Observatoire de Besançon for funding his visit, during which this paper was completed. He would also like to thank the direction and staff of Observatoire de Besançon for their hospitality. We acknowledge with thanks from the referee, Danielle Alloin, positive comments and constructive suggestions which have helped to improve the quality of this paper. For this research, we have made use of the Lyon-Meudon Extragalactic Database (LEDA), operated by the Lyon and Paris-Meudon Observatories (France). We also used the NASA/IPAC Extragalactic Database (NED) which is operated by the Jet Propulsion Laboratory, California Institute of Technology, under contract with the National Aeronautics and Space Administration.
warning/0001/cond-mat0001355.html
ar5iv
text
# Phase interference of spin tunneling in an arbitrarily directed magnetic field We present an exact analytic study on the topological phase interference effect in resonant quantum tunneling of the magnetization between degenerate excited levels for biaxial ferromagnets in an arbitrarily directed magnetic field. We show that the topological phase interference effect depends on the orientation of the field distinctly. The transition from classical to quantum behavior is also discussed. PACS number(s): 75.45.+j, 75.10.Jm, 03.65.Bz During the last decade, there has been great interest in the problem of quantum tunneling in nanometer-scale magnets. One notable subject is that the topological Wess-Zumino (or Berry) phase can lead to remarkable spin-parity effects. It was theoretically shown that the tunnel splitting is suppressed to zero for half-integer total spins in single-domain biaxial ferromagnetic (FM) particles due to the destructive interference of the Berry phase between two tunneling paths of opposite windings. However, the interference is constructive for integer spins, and hence the splitting is nonzero. While spin-parity effects are sometimes be related to the Kramers degeneracy, they typically go beyond the Kramers theorem in a rather unexpected way. The auxiliary particle method was proposed to study the spin-parity effects in one model of a single large spin subject to the external and anisotropy fields. Recently, the spin-phase interference and quantum oscillation effects were studied extensively in FM particles in the presence of a magnetic field, with the field along either the hard, easy, or medium axis. Similar spin-parity effects were also found in antiferromagnetic (AFM) particles, and in the quantum propagation of Bloch walls in quasi-one-dimensional ferromagnets and antiferromagnets. By applying the effective Hamiltonian approach, the effect of magnetic field and quantum interference of the magnetization vector (or the Néel vector) were studied in FM (or AFM) particles with different symmetry. Experiment on the molecules Fe<sub>8</sub> showed a direct evidence of the role of the topological phase in spin dynamics. Recent theoretical and experimental studies include the spin tunneling in a swept magnetic field, the thermally activated resonant tunneling based on the perturbation theory and the exact diagonalization, the discrete WKB method and a nonperturbation calculation, the non-adiabatic Landau-Zener model, the calculation based on exact spin-coordinate correspondence, and the effects caused by the higher order term and the nuclear spins on the tunnel splitting of Fe<sub>8</sub>. It is noted that the previous studies on FM spin-parity effects were mostly focused on the phase interference between two opposite winding ground-state tunneling paths. Moreover, the previous works have been confined to the condition that the magnetic field be applied along the easy, medium, or hard axis, separately. In this paper we study the topological phase interference effects for single-domain biaxial FM particles in an arbitrarily directed magnetic field. Our study provides a nontrivial generalization of the Kramers degeneracy for equivalent double-well system to coherently spin tunneling at ground states as well as low-lying excited states for FM system with asymmetric twin barriers caused by the arbitrarily directed magnetic field. Integrating out the momentum in the path integral, the spin tunneling problem is mapped onto a particle moving problem in one-dimensional periodic potential $`U\left(\varphi \right)`$. By applying the periodic instanton method, we obtain exactly the splittings between degenerate excited levels of neighboring wells. $`U\left(\varphi \right)`$ is regarded as a one-dimensional superlattice with a periodically recurring asymmetric twin barriers. The general translation symmetry results in the energy band structure, and the low-lying energy level spectrum is obtained by using the Bloch theorem. Our result shows that the excited-level splittings depend significantly on the parity of total spins of FM particles, and this spin-parity effect depends on the orientation of the field distinctly. The transition from quantum to classical behavior is also studied, and the second-order phase transition is shown. For a spin tunneling problem, the tunneling rate is determined by the Euclidean transition amplitude $$K_E=f\left|e^T\right|i=D\mathrm{\Omega }\mathrm{exp}\left(S_E\right),$$ (1) where $`D\mathrm{\Omega }=\mathrm{sin}\theta d\theta d\varphi `$, and the Euclidean action is $$S_E(\theta ,\varphi )=\frac{V}{\mathrm{}}𝑑\tau \left[i\frac{M_0}{\gamma }\left(\frac{d\varphi }{d\tau }\right)i\frac{M_0}{\gamma }\left(\frac{d\varphi }{d\tau }\right)\mathrm{cos}\theta +E(\theta ,\varphi )\right].$$ (2) $`M_0V=\left|\stackrel{}{M}\right|V=\mathrm{}\gamma S`$, where $`V`$ is the volume of the particle, $`\gamma `$ is the gyromagnetic ratio, and $`S`$ is the total spins. The first two terms in Eq. (2) define the topological Wess-Zumino term which arises from the nonorthogonality of spin coherent states. The total derivative has no effect on the classical equations of motion, but is crucial for the spin-parity effects. The system of interest has the biaxial symmetry, with $`\widehat{x}`$ being the easy axis, $`\widehat{y}`$ being the medium axis, and $`\widehat{z}`$ being the hard axis. The magnetic field is applied in the $`ZY`$ plane, at an angle in the range of $`\frac{\pi }{2}\theta _H\pi `$. Now $`E(\theta ,\varphi )`$ in Eq. (2) is $$E(\theta ,\varphi )=K_{}\mathrm{cos}^2\theta +K_{}\mathrm{sin}^2\theta \mathrm{sin}^2\varphi M_0H_y\mathrm{sin}\theta \mathrm{sin}\varphi M_0H_z\mathrm{cos}\theta ,$$ (3) where $`H_y=H\mathrm{sin}\theta _H`$, $`H_z=H\mathrm{cos}\theta _H`$, $`K_{}`$ and $`K_{}`$ are the longitudinal and the transverse anisotropy coefficients satisfying $`K_{}K_{}>0`$. As $`K_{}K_{}>0`$, the deviations of $`\theta `$ about $`\theta _0`$ are small. Introducing $`\theta =\theta _0+\alpha `$, $`\left|\alpha \right|1`$, Eq. (3) reduces to $$E(\alpha ,\varphi )=K_{}\mathrm{sin}^2\theta _0\alpha ^2+K_{}\mathrm{sin}^2\theta _0\left(\mathrm{sin}\varphi \mathrm{sin}\varphi _0\right)^2,$$ (4) where $`\mathrm{cos}\theta _0=\frac{M_0H_z}{2K_{}}`$, $`\mathrm{sin}\varphi _0=\frac{M_0H_y}{2K_{}\mathrm{sin}\theta _0}=\frac{h\mathrm{sin}\theta _H}{\sqrt{1\left(\lambda h\mathrm{cos}\theta _H\right)^2}}`$, $`\lambda =\frac{K_{}}{K_{}}`$, $`h=\frac{H}{H_0}`$, and $`H_0=\frac{2K_{}}{M_0}`$. Performing the Gaussian integration over $`\alpha `$, we obtain the transition amplitude as $$K_E=\mathrm{exp}\left[iS\left(1\mathrm{cos}\theta _0\right)\left(\varphi _f\varphi _i\right)\right]𝑑\varphi \mathrm{exp}\left\{𝑑\tau \left[\frac{1}{2}m\left(\frac{d\varphi }{d\tau }\right)^2+U\left(\varphi \right)\right]\right\},$$ (5) with $`m=\frac{\mathrm{}S^2}{2K_{}V}`$, and $`\mathrm{}U\left(\varphi \right)=K_{}V\mathrm{sin}^2\theta _0\left(\mathrm{sin}\varphi \mathrm{sin}\varphi _0\right)^2`$. Since the configuration space of this problem is a circle, calculations are restricted to the first twin barrier at $`\varphi =\frac{\pi }{2}`$ and $`\frac{3\pi }{2}`$. We use $`A`$ to denote the instanton passing through the small barrier at $`\varphi =\frac{\pi }{2}`$ with the height $`\mathrm{}U_S=K_{}V\mathrm{sin}^2\theta _0\left(1\mathrm{sin}\varphi _0\right)^2`$, and $`B`$ through the large barrier at $`\varphi =\frac{3\pi }{2}`$ with the height $`\mathrm{}U_L=K_{}V\mathrm{sin}^2\theta _0\left(1+\mathrm{sin}\varphi _0\right)^2`$. Correspondingly, there are two kinds of anti-instantons: $`A^{}`$ and $`B^{}`$. The periodic instanton configuration $`\varphi _p`$ which minimizes the Euclidean action in Eq. (5) at an energy $`E>0`$ satisfies the equation of motion $$\frac{1}{2}m\left(\frac{d\varphi _p}{d\tau }\right)^2U\left(\varphi _p\right)=E.$$ (6) Then we obtain the periodic instanton $`A`$ solution as $$\mathrm{sin}\varphi _A=\frac{1\xi _1\text{sn}^2(\omega _1\tau ,k_1)}{1+\xi _1\text{sn}^2(\omega _1\tau ,k_1)},$$ (7) corresponding to the transition of $`\varphi `$ from $`\mathrm{arcsin}\alpha `$ to $`\pi \mathrm{arcsin}\alpha `$, where $`\alpha =\mathrm{sin}\varphi _0+\sqrt{\frac{\mathrm{}E}{K_{}V\mathrm{sin}^2\theta _0}}`$. sn$`(\omega _1\tau ,k_1)`$ is the Jacobian elliptic sine function of modulus $`k_1`$, where $`k_1^2=\frac{\left(1\alpha \right)\left(1+\beta \right)}{\left(1+\alpha \right)\left(1\beta \right)}`$, $`\beta =\mathrm{sin}\varphi _0\sqrt{\frac{\mathrm{}E}{K_{}V\mathrm{sin}^2\theta _0}}`$, $`\xi _1=\frac{1\alpha }{1+\alpha }`$, $`\omega _1=\frac{\omega _0}{g_1}`$, $`\omega _0=2\frac{\sqrt{K_{}K_{}}V}{\mathrm{}S}\mathrm{sin}\theta _0`$, and $`g_1=\frac{2}{\sqrt{\left(1+\alpha \right)\left(1\beta \right)}}`$. The classical action or the WKB exponent in the tunnel splitting is obtained by integrating the Euclidean action in Eq. (5) with the above periodic instanton solution. The result for instanton $`A`$ is $`S_A=W_A+2E\beta `$, where $$W_A=4m\omega _1\left[E\left(k_1\right)+\frac{\left(k_1^2\xi _1\right)}{\xi _1}K\left(k_1\right)+\frac{\left(\xi _1^2k_1^2\right)}{\xi _1}\mathrm{\Pi }(k_1,\xi _1)\right].$$ (8) $`K\left(k_1\right)`$, $`E\left(k_1\right)`$, and $`\mathrm{\Pi }(k_1,\xi _1)`$ are the complete elliptic integral of the first, second, and third kind. Now we discuss briefly the calculation of the tunnel splitting. For a particle with mass $`m`$ moving in a smooth double-well-like one-dimensional potential $`U\left(x\right)`$, the instanton approach gives the ground-state tunnel splitting as $$\mathrm{\Delta }E_0=2C\left(\frac{W_0}{2\pi }\right)^{1/2}\mathrm{exp}\left(W_0\right),$$ (9) where $`W_0`$ is the classical action for the ground-state tunneling, $$C=\left\{\frac{det\left(_\tau ^2+\omega ^2\right)}{det^{}\left[_t^2+U^{\prime \prime }\left(x_{cl}\left(\tau \right)\right)/m\right]}\right\}^{1/2}$$ (10) is the ratio of fluctuation determinants, and $`x_{cl}\left(\tau \right)`$ is the instanton solution. The prime on the det indicates that the zero eigenvalue is to omitted, and $`\omega `$ is the frequency of harmonic amplitude oscillations in the wells about the minima. The general formulas were presented to evaluate the preexponential factor in the tunnel splitting or the tunneling rate. It was found that what is need for this evaluation is the asymptotic $`\left(\tau \pm \mathrm{}\right)`$ behavior of the instanton velocity, $$\frac{dx_{cl}}{d\tau }a\mathrm{exp}\left(\omega \tau \right),\text{ as }\tau \pm \mathrm{}.$$ (11) Then the ground-state tunnel splitting is $$\mathrm{\Delta }E_0=2\left|a\right|\left(\frac{m\omega }{\pi }\right)^{1/2}\mathrm{exp}\left(W_0\right).$$ (12) In most physical applications, however, the ground-state tunnel splitting can be best estimated as $$\mathrm{\Delta }E_0=p_0\omega \left(\frac{W_0}{2\pi }\right)^{1/2}\mathrm{exp}\left(W_0\right),$$ (13) where the dimensionless prefactor $`p_0`$ is often relevant to experiments. It is noted that Eq. (12) is based on quantum tunneling at the level of ground state, and the temperature dependence of the tunneling frequency (i.e., tunneling at excited levels) is not taken into account. The instanton technique is suitable only for the evaluation of the tunneling rate or the tunnel splitting at the vacuum level, since the usual (vacuum) instantons satisfy the vacuum boundary conditions. Recently, different types of pseudoparticle configurations (periodic or nonvacuum instantons) are found which satisfy periodic boundary conditions. For the same tunneling problem, the WKB answer for the splittings of the $`n`$th degenerate excited levels or the imaginary parts of the metastable levels is $$\mathrm{\Delta }E_n\left(\text{or }ImE_n\right)=\frac{\omega \left(E_n\right)}{\pi }\mathrm{exp}\left(W\right),$$ (14) where $`\omega \left(E_n\right)=\frac{2\pi }{t\left(E_n\right)}`$ is the energy-dependent frequency. $`t\left(E_n\right)`$ is the period of the real-time oscillation in the potential well, $$t\left(E_n\right)=\sqrt{2m}_{x_1\left(E_n\right)}^{x_2\left(E_n\right)}\frac{dx}{\sqrt{E_nU\left(x\right)}},$$ (15) where $`x_{1,2}\left(E_n\right)`$ are the turning points for the particle oscillating inside the potential $`U\left(x\right)`$. The functional-integral and the WKB method showed that for the potentials parabolic near the bottom the result Eq. (14) should be multiplied by $`\sqrt{\frac{\pi }{e}}\frac{\left(2n+1\right)^{n+1/2}}{2^ne^nn!}`$. This factor is very close to 1 for all $`n`$: 1.075 for $`n=0`$, 1.028 for $`n=1`$, 1.017 for $`n=2`$, etc. Stirling’s formula for $`n!`$ shows that this factor trends to 1 as $`n\mathrm{}`$. Therefore, this correction factor, however, does not change much in front of the exponentially small action term in Eq. (14). It is noted that Eq. (14) can be obtained by using the connection formulas near a linear turning point, or by matching the WKB wavefunction in the classically forbidden region to the exact harmonic oscillator wavefunction near the classical turning point. It was shown that the WKB method is equivalent to the instanton method for the ground-state tunnel splitting. Liang et al. showed that the WKB method and the periodic instanton method give the same result for the excited-state tunnel splitting. For the present problem, we find that $`\mathrm{\Delta }\epsilon _A=\frac{2}{t_A\left(E\right)}\mathrm{exp}\left(W_A\right)`$, where $`t_A\left(E\right)=\frac{2}{\omega _1\left(E\right)}K\left(k_1^{}\right)`$ and $`k_1^{}=\sqrt{1k_1^2}`$. The same method can be applied to the instanton $`B`$ passing through the large barrier. And the result is $`\mathrm{\Delta }\epsilon _B=\frac{2}{t_B\left(E\right)}\mathrm{exp}\left(W_B\right)`$ for $`0<E<U_S`$, where $`t_B\left(E\right)=t_A\left(E\right)`$, and $`W_B`$ has the same expression as Eq. (8) but taking $`\xi _1`$ as $`\xi _2=\frac{1+\beta }{1\beta }`$. For $`U_SEU_L`$, the imaginary part of the metastable level is $`ImE=\frac{2}{\stackrel{~}{t}_B\left(E\right)}\mathrm{exp}\left(2\stackrel{~}{W}_B\right)`$, where $`\stackrel{~}{t}_B\left(E\right)=\frac{2}{\omega _2\left(E\right)}K\left(k_2^{}\right)`$, $$\stackrel{~}{W}_B=2m\xi _3\left(1+\alpha \right)\omega _2\left[\frac{1}{k_2^2\xi _3}E\left(k_2\right)\frac{1}{\xi _3}K\left(k_2\right)+\frac{k_2^2+\xi _3^2+2k_2^2\xi _3}{\xi _3}\mathrm{\Pi }(k_2,\xi _3)\right],$$ (16) with $`k_2^{}=\sqrt{1k_2^2}`$, $`\omega _2=\frac{\omega _0}{g_2}`$, $`g_2=\sqrt{\frac{2}{\alpha \beta }}`$, $`\xi _3=\frac{1+\beta }{\alpha \beta }`$, and $`k_2^2=\frac{\left(\alpha 1\right)\left(1+\beta \right)}{2\left(\alpha \beta \right)}`$. Now we discuss the low energy limit of the level splitting. With the help of harmonic oscillator approximation for energy near the bottom of the potential well, $`\epsilon _n=\left(n+\frac{1}{2}\right)\mathrm{\Omega }`$, and $`\mathrm{\Omega }=\sqrt{{\displaystyle \frac{1}{m}}\left({\displaystyle \frac{d^2U}{d\varphi ^2}}\right)_{\varphi =\varphi _0}}=\sqrt{{\displaystyle \frac{2K_{}V}{\mathrm{}m}}}\mathrm{sin}\theta _0\mathrm{cos}\varphi _0,`$ $`W_{A\left(B\right)}`$ can be expanded as $`W_{A\left(B\right),n}`$ $`=`$ $`W_{A\left(B\right),0}\left(n+{\displaystyle \frac{1}{2}}\right)+\left(n+{\displaystyle \frac{1}{2}}\right)\mathrm{ln}\left({\displaystyle \frac{n+1/2}{8\sqrt{\lambda }S\mathrm{sin}\theta _0\mathrm{cos}^3\varphi _0}}\right),`$ () $`W_{A\left(B\right),0}`$ $`=`$ $`2\sqrt{\lambda }S\mathrm{sin}\theta _0\left(\mathrm{cos}\varphi _02\mathrm{sin}\varphi _0\mathrm{arcsin}\sqrt{{\displaystyle \frac{1\mathrm{sin}\varphi _0}{2}}}\right),`$ () where “$``$” for the instanton $`A`$, and “$`+`$” for the instanton $`B`$. Therefore, the tunnel splittings for instantons $`A`$ and $`B`$ are $$\mathrm{}\mathrm{\Delta }\epsilon _{A\left(B\right),n}=\frac{2^{3/2}}{\sqrt{\pi }n!}\left(K_{}V\right)\sqrt{\lambda }S^1\mathrm{sin}\theta _0\mathrm{cos}\varphi _0\left(8\sqrt{\lambda }S\mathrm{sin}\theta _0\mathrm{cos}^3\varphi _0\right)^{n+1/2}\mathrm{exp}\left(W_{A\left(B\right),0}\right).$$ (17) Equation (17b) shows that the WKB exponent for instanton $`A`$ is smaller than that for instanton $`B`$ at finite magnetic field because the barrier through which instanton $`B`$ must tunnel is higher than that for instanton $`A`$. For the case of ground-state resonance, i.e., $`n=0`$, Eq. (18) reduces to $$\mathrm{\Delta }\epsilon _{A\left(B\right),0}=\frac{4}{\sqrt{\pi }}\lambda ^{1/4}S^{1/2}\left(\mathrm{sin}\theta _0\mathrm{cos}^3\varphi _0\right)^{1/2}\mathrm{\Omega }\mathrm{exp}\left(W_{A\left(B\right),0}\right).$$ (18) It is not particularly illuminating to write out the prefactor $`p_0`$ (see Eq. (13)), although it can be seen that it is dimensionless and independent of the volume $`V`$ of the particle. Note that Eq. (13) is the approximate formula for ground-state tunnel splitting with exponential accuracy. In most physical applications, the preexponential factor can be best estimated as an attempt frequency. The apparent disagreement with Eq. (13) was also found in other spin-tunneling problems. However, in the case of zero magnetic field Eq. (19a) reduces to $`\mathrm{\Delta }\epsilon _{A,0}\left(H=0\right)`$ $`=`$ $`\mathrm{\Delta }\epsilon _{B,0}\left(H=0\right)`$ (19) $`=`$ $`4\mathrm{\Omega }_0\left({\displaystyle \frac{W_0}{2\pi }}\right)^{1/2}\mathrm{exp}\left(W_0\right),`$ () where $`\mathrm{\Omega }_0=\mathrm{\Omega }\left(H=0\right)`$, and $`W_0=W_{A,0}\left(H=0\right)=W_{B,0}\left(H=0\right)=2\sqrt{\lambda }S`$. Compared with Eq. (13), $`p_0=4`$ for this case. It is noted that $`\mathrm{}\mathrm{\Delta }\epsilon _{A\left(B\right),n}`$ is only the level shift induced by tunneling between degenerate excited states through a single barrier. $`U\left(\varphi \right)`$ can be regarded as a one-dimensional superlattice with the sublattices $`A`$ and $`B`$. The Bloch states for sublattices $`A`$ and $`B`$ are $$\mathrm{\Phi }_A(\xi ,\varphi )=\frac{1}{\sqrt{L}}\underset{n}{}e^{i\xi \varphi _n}\phi _A\left(\varphi \varphi _n\right),$$ (20) and $$\mathrm{\Phi }_B(\xi ,\varphi )=\frac{1}{\sqrt{L}}\underset{n}{}e^{i\xi \left(\varphi _n+a\right)}\phi _B\left(\varphi \varphi _na\right),$$ (21) where $`\varphi _n=2n\pi +\varphi _0`$, $`L=N\left(a+b\right)`$, $`a=\pi 2\varphi _0`$, and $`b=\pi +2\varphi _0`$. The total wavefunction is $$\mathrm{\Psi }_\xi \left(\varphi \right)=a_A\left(\xi \right)\mathrm{\Phi }_A(\xi ,\varphi )+a_B\left(\xi \right)\mathrm{\Phi }_B(\xi ,\varphi ).$$ (22) Including the Wess-Zumino phase, we derive the secular equation in the tight-binding approximation as $$\left[\begin{array}{cc}\epsilon _nE\left(\xi \right)\hfill & e^{i\left(\xi \mu \right)a}\mathrm{\Delta }\epsilon _{A,n}+e^{i\left(\xi \mu \right)b}\mathrm{\Delta }\epsilon _{B,n}\hfill \\ e^{i\left(\xi \mu \right)a}\mathrm{\Delta }\epsilon _{A,n}+e^{i\left(\xi \mu \right)b}\mathrm{\Delta }\epsilon _{B,n}\hfill & \epsilon _nE\left(\xi \right)\hfill \end{array}\right]\left[\begin{array}{c}a_A\left(\xi \right)\hfill \\ a_B\left(\xi \right)\hfill \end{array}\right]=0,$$ (23) where $`\mu =S\left(1\mathrm{cos}\theta _0\right)`$, and the Bloch wave vector $`\xi =0`$ in the first Brillouin zone. Then the tunnel splitting of the $`n`$th excited level is $$\mathrm{\Delta }\epsilon _n=2\sqrt{\left(\mathrm{\Delta }\epsilon _{A,n}\right)^2+\left(\mathrm{\Delta }\epsilon _{B,n}\right)^2+2\left(\mathrm{\Delta }\epsilon _{A,n}\right)\left(\mathrm{\Delta }\epsilon _{B,n}\right)\mathrm{cos}\left[2\pi S\left(1\mathrm{cos}\theta _0\right)\right]}.$$ (24) Another approach to obtain Eq. (23) is to calculate the transition amplitude directly. The subtle point in evaluating the transition amplitude is how to arrange the instantons and anti-instantons appropriately to satisfy the boundary condition. Note that one configuration starting from $`|\theta _0,\varphi _0`$ and ending at $`|\theta _0,\varphi _0`$ can be an arbitrary permutation of the pairs $`\left(AB\right)`$, $`\left(AA^{}\right)`$, $`\left(B^{}B\right)`$, and $`\left(B^{}A^{}\right)`$. Introducing $`s`$, $`t`$, $`p`$, and $`q`$ as the numbers of $`A`$, $`B`$, $`A^{}`$, and $`B^{}`$ in the instantons and anti-instantons pairs, and $`i`$, $`j`$, $`k`$, and $`l`$ as those of $`\left(AB\right)`$, $`\left(AA^{}\right)`$, $`\left(B^{}B\right)`$, and $`\left(B^{}A^{}\right)`$, we have $$i+j=s,i+k=t,j+l=p,k+l=q.$$ (25) Therefore, $`s+q=t+p`$, and only three variables are independent. Now the transition amplitude is $`K_E`$ $`=`$ $`\sqrt{{\displaystyle \frac{\mathrm{\Omega }}{\pi \mathrm{}}}}e^{\left(n+\frac{1}{2}\right)\mathrm{\Omega }}{\displaystyle \underset{s,t,p,q}{}}{\displaystyle \frac{N(s,t,p,q)}{\left(s+t+p+q\right)!}}[\left(\mathrm{}\mathrm{\Delta }\epsilon _{A,n}T\right)^{s+p}\left(\mathrm{}\mathrm{\Delta }\epsilon _{B,n}T\right)^{t+q}`$ () $`\times e^{iS\left(1\mathrm{cos}\theta _0\right)\left(\pi 2\varphi _0\right)\left(sp\right)}e^{iS\left(1\mathrm{cos}\theta _0\right)\left(\pi +2\varphi _0\right)\left(tq\right)}]\delta _{s+q,t+p},`$ where $`N(s,t,p,q)`$ represents the number of different configurations for a given set of $`\{s,t,p,q\}`$. It can be calculated as $$N(s,t,p,q)=\underset{i=\mathrm{max}\{0,tq\}}{\overset{\mathrm{min}\{s,t\}}{}}\frac{\left(i+j+k+l\right)!}{i!j!k!l!},$$ (27) which has a simple result $$N(s,t,p,q)=\frac{\left[\left(s+q\right)!\right]^2}{s!t!p!q!}.$$ (28) Then we have $$K_E=\sqrt{\frac{\mathrm{\Omega }}{\pi \mathrm{}}}e^{\left(n+\frac{1}{2}\right)\mathrm{\Omega }}\mathrm{cosh}\left[\sqrt{\left(\mathrm{\Delta }\epsilon _{A,n}\right)^2+\left(\mathrm{\Delta }\epsilon _{B,n}\right)^2+2\left(\mathrm{\Delta }\epsilon _{A,n}\right)\left(\mathrm{\Delta }\epsilon _{B,n}\right)\mathrm{cos}\left[2\pi S\left(1\mathrm{cos}\theta _0\right)\right]}T\right],$$ (29) and we can read off the splitting of the $`n`$th excited level shown in Eq. (23) directly. Finally we discuss the phase transition from classical to quantum behavior. It was found that for a particle moving in a double-well potential $`U\left(x\right)`$, the behavior of the energy-dependent period of oscillations $`P\left(E\right)`$ in the Euclidean potential $`U\left(x\right)`$ determines the order of the quantum-classical transition. If $`P\left(E\right)`$ monotonically increases with the amplitude of oscillations, i.e., with decreasing energy $`E`$, the transition is of second order. The corresponding crossover temperature is $`T_0^{\left(2\right)}=\frac{\stackrel{~}{\omega }_0}{2\pi }`$, $`\stackrel{~}{\omega }_0=\sqrt{\frac{1}{m}\left|U^{\prime \prime }\left(x_{sad}\right)\right|}`$ is the frequency of small oscillations near the bottom of $`U\left(x\right)`$, where $`x_{sad}`$ corresponds to the top (the saddle point) of the barrier. If, however, the dependence of $`P\left(E\right)`$ is non-monotonic, the first order crossover takes place. For this case, the period of the instanton $`A`$ is $`P_A\left(E\right)=\frac{4}{\omega _1}K\left(k_1\right)`$. The monotonically decreasing behavior of $`P_A\left(E\right)`$ in the domain $`0EU_S`$ is shown in Fig. 1 for several values of $`\mathrm{sin}\varphi _0`$, which shows that the second-order phase transition takes place, and the crossover temperature is $$k_BT_{0,A}^{\left(2\right)}=\frac{\sqrt{K_{}K_{}}V}{\pi \mathrm{}S}\mathrm{sin}\theta _0\sqrt{1\mathrm{sin}\varphi _0}.$$ (30) The period of the instanton $`B`$ is $`P_B\left(E\right)=P_A\left(E\right)`$ for $`0EU_S`$, while $`P_B\left(E\right)=\frac{4}{\omega _2}K\left(k_2\right)`$ for $`U_SEU_L`$. Fig. 2 shows the monotonically decreasing behavior of $`P_B\left(E\right)`$ for the whole domain $`0EU_L`$. Again one finds a second-order transition from the thermal to quantum regime, and the crossover temperature is $$k_BT_{0,B}^{\left(2\right)}=\frac{\sqrt{K_{}K_{}}V}{\pi \mathrm{}S}\mathrm{sin}\theta _0\sqrt{1+\mathrm{sin}\varphi _0}.$$ (31) At zero magnetic field, $`\mathrm{\Delta }\epsilon _{A,n}=\mathrm{\Delta }\epsilon _{B,n}`$, the tunnel splitting is suppressed to zero for the half-integer total spins by the destructive interfering Wess-Zumino-Berry phases, which is in good agreement with the Kramers theorem. The presence of a magnetic field perpendicular to the plane of rotation of magnetization yields an additional contribution to the Berry phase, resulting constructive and destructive interferences alternatively for both integer and half-integer spins. Tunneling is thus periodically suppressed. At finite magnetic field, the low-lying tunneling level spectrum depends on the parity of total spins. If $`S`$ is an integer, the tunnel splitting is $$\mathrm{\Delta }\epsilon _n=2\sqrt{\left(\mathrm{\Delta }\epsilon _{A,n}\right)^2+\left(\mathrm{\Delta }\epsilon _{B,n}\right)^2+2\left(\mathrm{\Delta }\epsilon _{A,n}\right)\left(\mathrm{\Delta }\epsilon _{B,n}\right)\mathrm{cos}\left(2\pi S\mathrm{cos}\theta _0\right)}.$$ (32) While if $`S`$ is a half-integer, the splitting is $$\mathrm{\Delta }\epsilon _n=2\sqrt{\left(\mathrm{\Delta }\epsilon _{A,n}\right)^2+\left(\mathrm{\Delta }\epsilon _{B,n}\right)^22\left(\mathrm{\Delta }\epsilon _{A,n}\right)\left(\mathrm{\Delta }\epsilon _{B,n}\right)\mathrm{cos}\left(2\pi S\mathrm{cos}\theta _0\right)}.$$ (33) Equations (30a) and (30b) clearly show oscillations of the tunnel splitting as a function of the transverse magnetic field, together with the parity effect. Note that the magnetic field along the medium axis does not produce any oscillations. As $`\mathrm{\Delta }\epsilon _{A,n}>\mathrm{\Delta }\epsilon _{B,n}`$ at finite magnetic field, the tunnel splitting will not be suppressed to zero even if the total spin is a half-integer. The similar oscillation of tunnel splitting with the magnetic field was also found in Ref. 29, while the system considered in Ref. 29 is a single-domain antiferromagnetic particle with a field along the hard axis. The most interesting observation is that only the $`\widehat{z}`$ component of the magnetic field (i.e., along the hard axis) can lead to the oscillation of the tunnel splitting. As shown in Fig. 3, the splitting depends on the magnitude of the field $`𝐇`$ and its angle $`\theta _H`$ with the hard axis, and even a small misalignment of the field with the $`\widehat{z}`$ axis can completely destroy the oscillation effect. For small $`\theta _H`$ the tunnel splitting oscillates with the field, whereas no oscillation is shown up for large $`\theta _H`$. In the latter case, a much stronger increase of tunnel splitting with the field is shown. This strong dependence on the orientation of the field can be observed for ground-state resonance as well as excited-state resonance. As a result, we conclude that the topological phase interference or spin-parity effects depend on the orientation of the external magnetic field distinctly. The distinct angular dependence, together with the oscillation of the tunnel splittings with the field, may provide an independent experimental test for the spin phase interference effects in FM particles. This “Aharonov-Bohm” type of oscillation in magnetic system is analogous to the oscillations as a function of external flux in a SQUID ring. Due to the topological nature of the Berry phase, these spin-parity effects are independent of details such as the magnitude of total spins and the shape of the soliton. The transition from classical to quantum behavior is also studied for the small and large tunneling barrier, respectively. By calculating the periods of instanton $`A`$ and $`B`$ analytically, we find the monotonically decreasing behavior of the periods with increasing energy, which yields a second-order phase transition. Our results should be useful for a quantitative understanding on the spin-parity effects and the quantum-classical crossover for FM particles in an arbitrarily directed magnetic field. It is noted that the theoretical results presented here are based on the instanton method, which is semiclassical in nature, i.e., valid for large spins and in the continuum limit. The theoretical calculations performed in this paper can be extended to the single-domain antiferromagnetic particles where the relevant quantity is the excess spin due to the small noncompensation of two sublattices. Work along this line is still in progress. We hope that the theoretical results obtained in the present work will stimulate more experiments whose aim is observing the topological phase interference effects and the quantum-classical transition in resonant quantum tunneling of magnetization in nanometer-scale single-domain ferromagnets.
warning/0001/math0001103.html
ar5iv
text
# An analysis on the shape equation for biconcave axisymmetric vesicles ## Introduction In this article, a vesicle is represented by a closed surface $`\mathrm{\Sigma }`$ in $`^3`$ with mean curvature $`H`$, surface area $`\left|\mathrm{\Sigma }\right|`$ and it encloses a volume $`V`$. Its geometric shape is modelled by minimizing a functional, sometimes called Helfrich functional, $$=_\mathrm{\Sigma }(2H+c_0)^2\mathrm{d}S+\lambda \left|\mathrm{\Sigma }\right|+pV$$ with some physical constant parameters $`c_0`$, $`\lambda `$, and $`p`$. The parameters carry the following meanings: $`c_0`$ is the spontaneous curvature, $`\lambda `$ is the tensile stress, and $`p=p_op_i`$ is the osmotic pressure difference between the outer ($`p_o`$) and inner ($`p_i`$) media. We take a sign convention that $`H`$ is negative for the standard sphere. It has been observed experimentally long ago that a red blood cell is of biconcave-discoid shape (which will be defined mathematically later). And it is a quest to find the appropriate theoretical model for the bending energy. Historically, the early model of Canham, \[C\], is purely geometric and it is equivalent to the Willmore functional, \[W, ch. 7\], which equals $``$ with $`c_0=\lambda =p=0`$, up to a constant. Certainly, from a differential geometer’s view, this cannot be the correct model because the unique minimum of the Willmore functional for topologically spherical vesicles is the round sphere. This is also observed by physicists \[DH\]. In fact, there are many important mathematical studies of the Willmore functional because of its geometric implications; for instance, the existence of minimizers among a certain topological class by Simon \[Si\], and the conformal properties by Li and Yau \[LY\]. We expect that their works may contribute to a certain extent to a deeper theoretical understanding of the energy $``$. Helfrich takes physical condition together with the Gaussian curvature into account and proposes a modified bending energy, \[H1\]. The shape of blood cells and some other biological membranes is closely related to the formation of lipid bilayer vesicles in aqueous medium (e.g. liquid crystal). The physical condition is based on the elasticity of lipid bilayer vesicles. According to the Gauss-Bonnet Theorem, the integral of the Gaussian curvature is a topological constant. Thus, within a certain topological class of $`\mathrm{\Sigma }`$, Helfrich’s bending energy can be reduced to $``$ above. Many properties of $``$ are yet to be discovered, though there are some experimental observations and numerical simulations, \[HDH, MB\]. The existence and uniqueness of its minimizer of a certain topology are still unknown. It is also not known whether the minimizer is symmetric in any sense. Answers to these questions require deep geometric analysis of the functional and studies in this direction are rare. Nevertheless, there are related works, such as, on similar functionals, \[Si, LY, Ni\]; or on surface flows, \[E\]. The Euler-Lagrange equation corresponding to $``$ is $$4\mathrm{}_\mathrm{\Sigma }H+2(2H+c_0)\left[2(H^2K)c_0H\right]2\lambda H+p=0.$$ In the past, much effort of physicists and biologists has been spent on studying axisymmetric solution to this variational equation, \[DH, L, Se, OH2, OY, NOO1, NOO2\]. By an axisymmetric surface, we mean an embedded surface $`\mathrm{\Sigma }`$ in $`^3`$ which is rotationally symmetric and has a reflection symmetry by the plane perpendicular to the rotational axis. With this additional assumption, the fourth order equation can be reduced to a second order one, usually referred to as the shape equation of axisymmetric vesicles. However, it is still unknown to scientists how the solution depends on the physical parameters and which parameters yield a solution corresponding to a biconcave surface. Our work is in this direction. We can derive conditions on the physical parameters for a solution having the biconcave shape. The conditions are easily expressed in terms of the cubic and quadratic polynomials denoted by $`Q(t)`$ $`=t^3+2c_0t^2+(c_0^2+\lambda )t{\displaystyle \frac{p}{2}};`$ $`R(t)`$ $`=Q(t)t^3.`$ Our main result can be stated as, ###### Theorem. For any $`c_0`$, $`\lambda `$, and $`p>0`$ such that every real root of $`Q`$ is positive, there exists axisymmetric biconcave surfaces which satisfies the Euler-Lagrange equation of $``$. It should be remarked that when $`c_0>0`$, $`\lambda >0`$, and $`p>0`$, the condition is always satisfied. Helfrich’s numercial simulation produces a biconcave shape resembling a blood cell when $`c_0`$ is positive. Furthermore, we are also able to numerically construct other interesting shapes when the condition is not satisfied. One is multiconcave and the other has no reflection symmetry. This article is organized in the following way. In §1, we first give the differential equation for the revolving graph of an axisymmetric solution. The derivation of the equation is given in the appendix. Moreover, we formulate the problem of finding special solution corresponding to a biconcave shape surface. We also present several variants of the equation which will be useful later. The variants of the shape equation are then studied in §2 to show that our condition stated above is sufficient for getting the expected special solution for the problem. The analysis and the estimates of geometric quantities are discussed in detail here. Finally, since the solution and its reflection no longer form a graph at the reflection plane. We will show that the solution obtained in previous sections is still a solution across the reflection symmetry. It is sufficient to verify that it also satisfies the variation equation at the reflection plane. In the process, one more necessary geometric condition is obtained. ### Acknowledgement We would like to thank our colleague K. S. Chou, who has been encouraging in our project and making valuable suggestions. ## 1. The Equation, the Problem, and the Conditions An axisymmetric surface is a closed embedded surface $`\mathrm{\Sigma }`$ in $`^3`$ with a rotation symmetry and a reflection symmetry by the plane perpendicular to the rotation axis. It is biconcave if there are exactly two components of negative Gaussian curvature. Without loss of generality, the rotational axis is labelled the $`z`$-axis and the plane of reflection is the $`xy`$-plane. Then the surface $`\mathrm{\Sigma }`$ can be obtained by revolving a radial curve about the $`z`$-axis on the upper half plane and reflecting it to the lower half. Typically, a biconcave one is obtained by revolving and reflecting a curve shown in the picture below. A cross-section of a biconcave axisymmetric surface (with $`c_0=1`$, $`\lambda =0.25`$, $`p=1`$) In other words, one may parametrize the upper part of $`\mathrm{\Sigma }`$ by $`𝐗=(r\mathrm{cos}\theta ,r\mathrm{sin}\theta ,z(r)),`$ with a function $`z(r)`$ defined for $`r`$ in some interval $`[0,r_{\mathrm{}}]`$, where $`z(r)>0`$ for $`r[0,r_{\mathrm{}})`$ and $`z(r_{\mathrm{}})=0`$. There are natural boundary conditions imposed by the rotation and reflection symmetries of the surface. The obvious ones are $`z^{}(0)=0`$ and $`z^{}(r)\mathrm{}`$ as $`rr_{\mathrm{}}`$. However, more subtle ones arise from the regularity at the end-point $`r_{\mathrm{}}`$, which we will discuss later in §3. For a biconcave surface, there is $`r_M<r_{\mathrm{}}`$ such that $`z^{\prime \prime }(r)>0`$ if and only if $`r[0,r_M)`$. This is equivalent to require that $`z^{}`$ has only a unique maximum and no other critical point. For a biconcave axisymmetric surface, the variation equation for the Helfrich functional is reduced to a second order ordinary differential equation, traditionally called the shape equation. With the notation $`w(r)=z^{}(r)`$, the shape equation is (1) $`{\displaystyle \frac{2r}{(1+w^2)^{5/2}}}w^{\prime \prime }=`$ $`{\displaystyle \frac{5rw}{(1+w^2)^{7/2}}}w_{}^{}{}_{}{}^{2}{\displaystyle \frac{2w^{}}{(1+w^2)^{5/2}}}`$ $`+{\displaystyle \frac{2w+w^3}{r(1+w^2)^{3/2}}}+{\displaystyle \frac{2c_0w^2}{1+w^2}}+{\displaystyle \frac{(c_0^2+\lambda )rw}{(1+w^2)^{1/2}}}{\displaystyle \frac{pr^2}{2}}.`$ We leave the derivation of this equation to the appendix to focus on the idea and analysis of the equation. We are going to study special solution $`w(r)`$ to the initial value problem on this variational equation with initial choice $`w(0)=0`$ and $`w^{}(0)=w_0^{}>0`$. Specifically, we look for a solution with the additional requirements that it is unimodal; and there is a finite number $`r_{\mathrm{}}`$ such that $`w(r)\mathrm{}`$ as $`rr_{\mathrm{}}`$; and $`\mathrm{}<{\displaystyle _0^r_{\mathrm{}}}w(r)\mathrm{d}r<0.`$ In \[DH, JS, ZL\], there are versions of the same equation written in term of the angle $`\psi `$ between the surface tangent and the plane perpendicular to rotational axis. However, it is convenient for our discussion to write the equation in the above form. With our notation of the polynomials $`Q`$ and $`R`$, after multiplying with $`rw^{}`$, the equation is equivalent to, (1a) $`\left[{\displaystyle \frac{r^2w_{}^{}{}_{}{}^{2}}{(1+w^2)^{5/2}}}\right]^{}`$ $`=\left[{\displaystyle \frac{w^2}{(1+w^2)^{1/2}}}\right]^{}+r^3w^{}R(\kappa (r));\text{or}`$ (1b) $`\left[{\displaystyle \frac{r^2w_{}^{}{}_{}{}^{2}}{(1+w^2)^{5/2}}}\right]^{}`$ $`=\left[{\displaystyle \frac{2}{\sqrt{1+w^2}}}\right]^{}+r^3w^{}Q(\kappa (r)),`$ where $`\kappa (r)={\displaystyle \frac{w}{r\sqrt{1+w^2}}}`$. It will be seen that these groupings of the lower order terms are important in the analysis the equation. The understanding of $`\kappa (r)`$ also provides useful information about the solution. First, its derivatives are given by, $`\kappa ^{}(r)`$ $`={\displaystyle \frac{w^{}}{r(1+w^2)^{3/2}}}{\displaystyle \frac{w}{r^2\sqrt{1+w^2}}},`$ $`\kappa ^{\prime \prime }(r)`$ $`={\displaystyle \frac{w^{\prime \prime }}{r(1+w^2)^{3/2}}}{\displaystyle \frac{3ww_{}^{}{}_{}{}^{2}}{r(1+w^2)^{5/2}}}{\displaystyle \frac{2w^{}}{r^2(1+w^2)^{3/2}}}+{\displaystyle \frac{2w}{r^3\sqrt{1+w^2}}}.`$ Then we have the equation for $`\kappa (r)`$, (2) $$r\kappa ^{\prime \prime }=\frac{r\kappa (r\kappa ^{}+\kappa )^2}{2(1r^2\kappa ^2)}3\kappa ^{}+\frac{rQ(\kappa (r))}{2(1r^2\kappa ^2)}.$$ Let us end this section by remarking on a few geometric quantities in terms of $`w`$. Firstly, $`\kappa ={\displaystyle \frac{w}{r\sqrt{1+w^2}}}`$ is the principle curvature in the meridinal (rotational) direction. Another principle curvature, the longitudinal one, is given by $`{\displaystyle \frac{w^{}}{(1+w^2)^{3/2}}}`$, which occurs in $`\kappa ^{}`$. Their product is the Gaussian curvature, which is expected to be positive at the rotational axis and reflection plane but negative somewhere along the circle defined by the zero of $`w`$. ## 2. Analysis of the Equation In this section, we prove the conditions for the existence of the required special solution to the initial value problem on equation (1) described in the preceding section. The strategy is the following analysis on the equation. We study the principle curvature $`\kappa `$, which is positive initially (at the rotational axis), i.e., $`\kappa (0)=w_0^{}>0`$. If $`w_0^{}`$ is not too large in terms of the roots of $`Q`$, $`\kappa `$ must decrease and eventually becomes zero at $`r_0`$ for some $`0<r_0<r_{\mathrm{}}`$. After $`w`$ becomes negative, it continues its descent and blows down to $`\mathrm{}`$ at finite distance $`r_{\mathrm{}}`$. This guarantees C1 that $`w`$ is unimodal and has a unique zero at $`r_0`$. In order to verify that the solution satisfies the requirements C2 and C3, we establish, in terms of $`w_0^{}`$, the estimates on $`r_0`$, $`r_{\mathrm{}}`$ and values of $`w`$, $`w^{}`$ at these positions. ###### Lemma 2.1. Let $`w_0^{}>0`$ and $`R(t)<0`$ for $`t[0,w_0^{}]`$. If $`w0`$ on an interval $`[0,r_0)`$, then $`\kappa (r)`$ decreases on $`(0,r_0)`$. ###### Proof. By continuity of $`\kappa `$ and that $`\underset{r0}{lim}\kappa (r)=w_0^{}>0`$, there exists $`\epsilon >0`$ such that $`R(\kappa (r))<0`$, $`w0`$, and $`w^{}>0`$ on $`(0,\epsilon )`$. Integration of the equation (1a) on $`(0,\epsilon )`$ gives $$\frac{r^2w_{}^{}{}_{}{}^{2}}{(1+w^2)^{5/2}}<\frac{w^2}{\sqrt{1+w^2}}\text{on}(0,\epsilon ).$$ Thus, $`\kappa ^{}(r)={\displaystyle \frac{w^{}}{r(1+w^2)^{3/2}}}{\displaystyle \frac{w}{r^2\sqrt{1+w^2}}}<0`$ for $`r(0,\epsilon )`$. This also implies that $`\kappa (r)`$ remains in the interval $`(0,w_0^{}]`$ for $`r(0,\epsilon )`$ and hence the argument works before we hit the first zero $`r_M`$ of $`w^{}`$, i.e., the first critical point of $`w`$. This shows that $`r_1=inf\{r(0,r_0):\kappa ^{}(r)0\}r_M>0`$ if the set is nonempty. Now by the smoothness of $`\kappa `$, we conclude that $`\kappa (r_1)[0,w_0^{}]`$, $`\kappa ^{}(r_1)=0`$ and $`\kappa ^{\prime \prime }(r_1)0`$. Putting this into the equation (2), we have $$0r_1\kappa ^{\prime \prime }(r_1)=\frac{r_1R(\kappa (r_1))}{2\left(1r_1^2\kappa ^2(r_1)\right)}<0.$$ This is a contradiction and hence the set $`\{r(0,r_0):\kappa ^{}(r)0\}`$ must be empty. $`\mathrm{}`$ By the same token, we can also establish a criterion for a growing solution. Since we don’t need it for our further discussion, we will omit the proof. ###### Proposition 2.2. If $`c_0>0`$, $`p>0`$, $`w_0^{}>0`$ and $`R(w_0^{})>0`$, then $`w`$ is increasing and blows-up to $`+\mathrm{}`$. For our future discussion, we denote a few quantities which depends only on the polynomial $`Q`$ and the initial data $`w_0^{}`$ as follow: $`\mu (w_0^{})`$ $`=\mathrm{min}\left\{Q(t):0tw_0^{}\right\}=\mathrm{max}\left\{Q(t):0tw_0^{}\right\};`$ $`\delta _+(w_0^{})`$ $`=\mathrm{min}\left\{Q(t):0tw_0^{}\right\}=\mathrm{max}\left\{Q(t):0tw_0^{}\right\};`$ $`\delta _{}`$ $`=\mathrm{min}\left\{Q(t):t0\right\}=\mathrm{max}\left\{Q(t):t0\right\}.`$ Note that $`\underset{w_0^{}0}{lim}\delta _+(w_0^{})=p/2=\underset{w_0^{}0}{lim}\mu (w_0^{})`$, and $`\delta _{}`$ is independent of $`w_0^{}`$. We will now show that under reasonable condition, the solution will not blow up to $`+\mathrm{}`$. It is because the cubic lower terms are dominated by $`\delta _+`$. ###### Lemma 2.3. Suppose that all real roots of $`Q`$ are positive. If $`Q<0`$ on $`[0,w_0^{}]`$, i.e., $`w_0^{}`$ less than the smallest real root of $`Q`$, then there is $`r_0>0`$ with $`r_0^2<\frac{16w_0^{}}{\delta _+}`$ such that on the interval $`(0,r_0)`$, $$w>0,w(r_0)=0,\text{and}w^{}(r_0)<\frac{\delta _+}{8}r_0^2.$$ In addition, if $`64w_{0}^{}{}_{}{}^{3}<27\delta _+`$, then $$_0^{r_0}w(r)\mathrm{d}r\frac{4w_{0}^{}{}_{}{}^{2}}{\delta _+^{3/2}}\left(1\frac{64w_{0}^{}{}_{}{}^{3}}{27\delta _+}\right)^{1/2}.$$ ###### Proof. Consider the principal curvature $`\kappa (r)={\displaystyle \frac{w}{r\sqrt{1+w^2}}}`$ and rewrite equation (1) or (1a) as $$\kappa ^{\prime \prime }=\frac{3\kappa ^{}}{r}\frac{ww_{}^{}{}_{}{}^{2}}{r(1+w^2)^{5/2}}+\frac{1+w^2}{2}Q(\kappa (r)).$$ By lemma 2.1, $`\kappa ^{}(r)<0`$ on a neighborhood of $`0`$ and hence $`\kappa (r)<w_0^{}`$ near $`0`$. Using the assumption on $`Q`$ and the positivity of $`w`$ near $`0`$, we have $$r^3\kappa ^{\prime \prime }+3r^2\kappa ^{}\frac{\delta _+}{2}r^3.$$ Therefore, for small $`r`$, (3) $$\kappa ^{}(r)\frac{\delta _+}{8}r^2,$$ and (4) $$\kappa (r)w_0^{}\frac{\delta _+}{16}r^2.$$ These inequalities show that $`\kappa `$ remains in $`(0,w_0^{})`$ as long as $`w0`$. In turns, they themselves hold as long as $`w0`$. Clearly, they guarantee the existence of $`r_0`$ with $`r_0^2{\displaystyle \frac{16w_0^{}}{\delta _+}}`$ such that $`\kappa (r_0)=0`$ and so $`w(r_0)=0`$. Then by (3), we have $`w^{}(r_0)<\delta _+r_0^2/8`$. This proves the first statement of the theorem. To prove the second statement, we temporarily let $`a=w_0^{}`$ and $`b=\delta _+/16`$. Then on $`[0,r_0)`$, $$\frac{w}{r\sqrt{1+w^2}}abr^2.$$ With the assumption that $`4a^3<27b`$, it can be written as $$w\frac{r(abr^2)}{\sqrt{1r^2(abr^2)^2}}.$$ Therefore, $$w\frac{r(abr^2)}{\sqrt{1\frac{4a^3}{27b}}},$$ and $`{\displaystyle _0^{r_0}}w\mathrm{d}r`$ $`{\displaystyle _0^{\sqrt{a/b}}}{\displaystyle \frac{r(abr^2)}{\sqrt{1\frac{4a^3}{27b}}}}\mathrm{d}r`$ $`{\displaystyle \frac{1}{\sqrt{b\frac{4a^3}{27}}}}{\displaystyle \frac{a^2}{4b}}={\displaystyle \frac{4w_{0}^{}{}_{}{}^{2}}{\delta _+}}\left(1{\displaystyle \frac{64w_{0}^{}{}_{}{}^{3}}{27\delta _+}}\right)^{1/2}.`$ This completes the proof of the lemma. $`\mathrm{}`$ Immediately from the lemma we have the following ###### Corollary 2.4. Suppose that all real roots of $`Q`$ are positive. Then $$\underset{w_0^{}0}{lim\; sup}\frac{r_0^2}{w_0^{}}\frac{32}{p}\text{and}\underset{w_0^{}0}{lim\; sup}\frac{1}{w_{0}^{}{}_{}{}^{2}}_0^{r_0}w\mathrm{d}r\frac{8}{p}.$$ In the above, we obtained an upper bound for $`r_0`$, where $`w`$ first hits zero. Next, a lower bound is established, which is essential for future estimates. We note that from lemma 2.3, $`w`$ has at least one maximum in the interval $`(0,r_0)`$. Since $`w`$ is increasing at the beginning, the first critical point must be a maximum point. From now on, we let $`\xi =\xi (w_0^{})=1{\displaystyle \frac{64w_{0}^{}{}_{}{}^{3}}{27\delta _+}}>0`$. ###### Lemma 2.5. Let $`r_M(0,r_0)`$ be the first critical point of $`w`$. Then $$\underset{w_0^{}0}{lim\; inf}\frac{r_0^2}{w_0^{}}\underset{w_0^{}0}{lim\; inf}\frac{r_M^2}{w_0^{}}\frac{32}{3p},\underset{w_0^{}0}{lim\; inf}\frac{w^{}(r_0)}{w_0^{}}2.$$ ###### Proof. Using equation (4), $`\kappa >0`$ on $`[0,r_0)`$, and $`r_0^2<16w_0^{}/\delta _+`$, one can check that $$1r^2\kappa (r)^2\xi \text{on }(0,r_0).$$ Putting together the assumption on $`Q`$, the fact that $`\kappa <w_0^{}`$ on $`(0,r_0)`$, and the inequality (3) into the equations (2), one obtains $`\left(r\kappa \right)^{\prime \prime }`$ $`{\displaystyle \frac{1}{2\xi }}r\kappa (r\kappa ^{}+\kappa )^2+{\displaystyle \frac{1}{8}}\delta _+r{\displaystyle \frac{\mu }{2\xi }}r.`$ $`{\displaystyle \frac{1}{2\xi }}rw_0^{}\left[(r\kappa )^{}\right]^2+{\displaystyle \frac{1}{8}}\delta _+r{\displaystyle \frac{\mu }{2\xi }}r`$ $`={\displaystyle \frac{w_0^{}}{2\xi }}\left[(r\kappa )_{}^{}{}_{}{}^{2}+{\displaystyle \frac{\mu \frac{\xi }{4}\delta _+}{w_0^{}}}\right]r.`$ Note that in the above, $`\mu =\mathrm{max}\left\{Q(t):0tw_0^{}\right\}\delta _+`$. Let $`y=(r\kappa )^{}={\displaystyle \frac{w^{}}{(1+w^2)^{3/2}}}`$ and $`A^2={\displaystyle \frac{\mu \frac{\xi }{4}\delta _+}{w_0^{}}}{\displaystyle \frac{\left(1\frac{\xi }{4}\right)\delta _+}{w_0^{}}}>{\displaystyle \frac{3\delta _+}{4w_0^{}}}>0`$. We then have $$\frac{y^{}}{y^2+A^2}\frac{w_0^{}}{2\xi }r.$$ Integrating from $`0`$, it yields (5) $$\mathrm{arctan}\left(\frac{y}{A}\right)\mathrm{arctan}\left(\frac{w_0^{}}{A}\right)\frac{Aw_0^{}}{4\xi }r^2.$$ Taking $`r=r_M`$ and multiplying by $`A/w_0^{}`$, as $`y(r_M)=0`$ and $`r_0r_M`$, we have $`A^2r_0^2A^2r_M^2`$ $`4\xi {\displaystyle \frac{A}{w_0^{}}}\mathrm{arctan}\left({\displaystyle \frac{w_0^{}}{A}}\right)4\xi \left(1{\displaystyle \frac{w_{0}^{}{}_{}{}^{2}}{3A^2}}\right).`$ Then, $`{\displaystyle \frac{r_0^2}{w_0^{}}}{\displaystyle \frac{r_M^2}{w_0^{}}}`$ $`{\displaystyle \frac{4\xi }{\left(\mu \frac{\xi }{4}\delta _+\right)}}\left[1{\displaystyle \frac{w_{0}^{}{}_{}{}^{3}}{3\left(\mu \frac{\xi }{4}\delta _+\right)}}\right].`$ Considering the limiting situation, by $`\underset{w_0^{}0}{lim}\mu =\underset{w_0^{}0}{lim}\delta _+=p/2`$ and $`\underset{w_0^{}0}{lim}\xi =1`$, the first result follows. For the second estimate, we first take $`rr_0`$ in (5), apply $`r_0^2<16w_0^{}/\delta _+`$ to obtain $`{\displaystyle \frac{A(w^{}(r_0)w_0^{})}{A^2+w^{}(r_0)w_0^{}}}`$ $`\mathrm{tan}\left({\displaystyle \frac{4Aw_{0}^{}{}_{}{}^{2}}{\delta _+\xi }}\right){\displaystyle \frac{4Aw_{0}^{}{}_{}{}^{2}}{\delta _+\xi }}\left[\mathrm{cos}\left({\displaystyle \frac{4Aw_{0}^{}{}_{}{}^{2}}{\delta _+\xi }}\right)\right]^1.`$ Then, $`{\displaystyle \frac{w^{}(r_0)}{w_0^{}}}`$ $`{\displaystyle \frac{\mathrm{cos}\left(\frac{4Aw_{0}^{}{}_{}{}^{2}}{\delta _+\xi }\right)\frac{4A^2w_0^{}}{\delta _+\xi }}{\mathrm{cos}\left(\frac{4Aw_{0}^{}{}_{}{}^{2}}{\delta _+\xi }\right)+\frac{4w_{0}^{}{}_{}{}^{3}}{\delta _+\xi }}}.`$ This gives the second estimate immediately by letting $`w_0^{}0`$. $`\mathrm{}`$ Combining the previous lemmas, we have the following ###### Corollary 2.6. For any $`\epsilon >0`$, there exists $`\beta >0`$ such that if $`w_0^{}(0,\beta )`$, then $$\frac{32}{3p}\epsilon \frac{r_M^2}{w_0^{}}\frac{r_0^2}{w_0^{}}\frac{32}{p}+\epsilon $$ and $$2\epsilon \frac{w^{}(r_0)}{w_0^{}}\frac{2}{3}+\epsilon $$ The $`r_M`$ that we discussed in the previous lemma is in fact the unique critical point of $`w`$ in $`(0,r_0)`$ provided $`w_0^{}`$ is sufficiently small. ###### Lemma 2.7. Under conditions of lemmas 2.3 and 2.5, if $`w_0^{}`$ is sufficiently small, then $`w`$ has a unique maximum at $`r_M[0,r_0)`$ and has no other critical point. ###### Proof. Suppose $`w`$ has another critical points in $`(0,r_0)`$, then according to the definition of $`r_M`$, $`w`$ attains a positive nonmaximum critical at $`r_m(r_M,r_0)`$, i.e., $`w^{}(r_m)=0`$ and $`w^{\prime \prime }(r_m)0`$. By the corollary 2.6, given any $`\epsilon >0`$, $`r_m^2>r_M^2\left(\frac{32}{3p}\epsilon \right)w_0^{}`$ for sufficiently small $`w_0^{}`$. On the other hand, substituting $`w^{}(r_m)=0`$ and $`w^{\prime \prime }(r_m)0`$ into the equation (1), we have $`0`$ $`{\displaystyle \frac{2w_m}{r_m^3(1+w_{m}^{}{}_{}{}^{2})^{3/2}}}+Q(\kappa (r_m))`$ $`{\displaystyle \frac{2}{r_m^2}}\kappa (r_m)+Q(\kappa (r_m)),`$ where $`w_m=w(r_m)`$. Since $`0<\kappa (r_m)w_0^{}`$, thus, for sufficiently small $`w_0^{}`$, we also have $`Q(\kappa (r_m))<p/2+\epsilon `$. This leads to $$0\frac{2}{\frac{32}{3p}\epsilon }\frac{p}{2}+\epsilon $$ which is clearly a contradiction since $`\epsilon `$ is arbitrary. $`\mathrm{}`$ Next, we claim that under our condition on $`Q`$, after hitting $`0`$, $`w`$ decreases and goes to negative infinity in finite distance. ###### Theorem 2.8. Suppose that all real roots of $`Q(t)`$ are positive and $`r_0>0`$ is given by lemma 2.3. Then $`w^{}<0`$ for $`rr_0`$ and there is a finite number $`r_{\mathrm{}}>r_0`$ such that $`\underset{rr_{\mathrm{}}}{lim}w(r)=\mathrm{}`$. Moreover, we have $$r_{\mathrm{}}r_0\frac{\pi }{2\sqrt{\delta r_0^2\left|w^{}(r_0)\right|}},$$ where $`\delta =\mathrm{min}\{\frac{\delta _+}{8},\frac{\delta _{}}{2}\}`$ and $`\delta _{}=\mathrm{min}\left\{Q(t):t0\right\}`$. ###### Proof. For convenience, let us temporarily denote $`v=w`$ for $`r>r_0`$. Since $`\kappa (r)<0`$ and $`v^{}>0`$ on $`(r_0,r_0+\epsilon )`$ for some $`\epsilon >0`$, we have $`v^{}Q(\kappa (r))>\delta _{}v^{}`$ on the same interval and the equation (1b) gives (6) $$\left[\frac{r^2v_{}^{}{}_{}{}^{2}}{(1+v^2)^{5/2}}\right]^{}=\left[\frac{2}{(1+v^2)^{1/2}}\right]^{}r^3v^{}Q(\kappa (r))\left[\frac{2}{(1+v^2)^{1/2}}\right]^{}+\delta _{}r^3v^{}$$ on $`(r_0,r_0+\epsilon )`$. In particular, $$\left[\frac{r^2v_{}^{}{}_{}{}^{2}}{(1+v^2)^{5/2}}\right]^{}\left[\frac{2}{(1+v^2)^{1/2}}\right]^{},$$ which gives $$\frac{r^2v_{}^{}{}_{}{}^{2}}{(1+v^2)^{5/2}}r_0^2v^{}(r_0)^2\frac{2}{(1+v^2)^{1/2}}+20.$$ This implies that $`v^{}`$ does not vanish and therefore (7) $$rv^{}\frac{rv^{}}{(1+v^2)^{5/4}}r_0v^{}(r_0)>0.$$ This, in turns, implies that $`v`$ is increasing, $`\kappa `$ remains negative and (6) holds as long as $`v`$ is defined. Substitute (7) into the equation (6) again, it gives $$\left[\frac{r^2v_{}^{}{}_{}{}^{2}}{(1+v^2)^{5/2}}\right]^{}\delta _{}r^3v^{}\delta _{}r_0v^{}(r_0)r^2\delta _{}r_0^2v^{}(r_0)r$$ for all $`r>r_0`$ such that $`v`$ is defined. Now, integrating the above from $`r_0`$, we conclude that $$\frac{r^2v_{}^{}{}_{}{}^{2}}{(1+v^2)^{5/2}}r_0^2v^{}(r_0)^2+\frac{\delta _{}r_0^2v^{}(r_0)}{2}(r^2r_0^2).$$ In the case that $`\delta _+4\delta _{}`$, we have $`v^{}(r_0)\delta _{}r_0^2/2`$ by lemma 2.3 $`{\displaystyle \frac{r^2v_{}^{}{}_{}{}^{2}}{(1+v^2)^{5/2}}}`$ $`r_0^2v^{}(r_0)\left(v^{}(r_0){\displaystyle \frac{\delta _{}}{2}}r_0^2\right)+{\displaystyle \frac{\delta _{}r_0^2v^{}(r_0)}{2}}r^2`$ $`{\displaystyle \frac{\delta _{}r_0^2v^{}(r_0)}{2}}r^2.`$ Otherwise, for $`\zeta =\frac{\delta _+}{4\delta _{}}<1`$, one has $`r_0^2v^{}(r_0)^2\frac{1}{2}\zeta \delta _{}v^{}(r_0)r_0^4.`$ It follows that $`{\displaystyle \frac{r^2v_{}^{}{}_{}{}^{2}}{(1+v^2)^{5/2}}}`$ $`r_0^2v^{}(r_0)^2+{\displaystyle \frac{\delta _{}r_0^2v^{}(r_0)}{2}}(r^2r_0^2)`$ $`=r_0^2v^{}(r_0)^2+{\displaystyle \frac{\zeta \delta _{}r_0^2v^{}(r_0)}{2}}(r^2r_0^2)+{\displaystyle \frac{(1\zeta )\delta _{}r_0^2v^{}(r_0)}{2}}(r^2r_0^2)`$ $`r_0^2v^{}(r_0)\left[v^{}(r_0){\displaystyle \frac{\zeta \delta _{}v^{}(r_0)}{2}}r_0^2\right]+{\displaystyle \frac{\zeta \delta _{}r_0^2v^{}(r_0)}{2}}r^2.`$ Therefore, $`{\displaystyle \frac{r^2v_{}^{}{}_{}{}^{2}}{(1+v^2)^2}}`$ $`{\displaystyle \frac{r^2v_{}^{}{}_{}{}^{2}}{(1+v^2)^{5/2}}}{\displaystyle \frac{\zeta \delta _{}r_0^2v^{}(r_0)}{2}}r^2={\displaystyle \frac{1}{8}}\delta _+r_0^2v^{}(r_0)r^2.`$ Hence, with $`\delta =\mathrm{min}\{\frac{1}{4}\delta _+,\frac{1}{2}\delta _{}\},`$ one has $$\frac{v^{}}{1+v^2}\frac{v^{}}{(1+v^2)^{5/4}}\sqrt{\delta r_0^2v^{}(r_0)}$$ as long as $`v`$ is defined. Further integrating from $`r_0`$ gives $$\mathrm{arctan}v\sqrt{\delta r_0^2v^{}(r_0)}(rr_0).$$ This clearly shows that $`v`$ blows up before $`r`$ goes to infinity. Furthermore, the upper bound of $`r_{\mathrm{}}`$ follows. $`\mathrm{}`$ ###### Corollary 2.9. Under the same conditions of theorem 2.8, we have, $$\mathrm{}>_{r_0}^r_{\mathrm{}}\left|w(r)\right|\mathrm{d}rB^1\mathrm{log}\left(\frac{1}{\mathrm{cos}(B(r_{\mathrm{}}r_0))}\right)\frac{B}{2}(r_{\mathrm{}}r_0)^2,$$ where $`B=\sqrt{\delta r_0^2v^{}(r_0)}\sqrt{\frac{\delta \delta _+}{8}}r_0^2\frac{1}{\sqrt{2}}\delta r_0^2`$. ###### Proof. From (6), by simply dropping the term $`r^3v^{}Q(\kappa (r))>0`$, we have $$\frac{r^2v_{}^{}{}_{}{}^{2}}{(1+v^2)^{5/2}}2\left(1\frac{1}{\sqrt{1+v^2}}\right).$$ After multiplying by $`\sqrt{1+v^2}(\sqrt{1+v^2}+1)`$, it becomes $`(\sqrt{1+v^2}+1){\displaystyle \frac{r^2v_{}^{}{}_{}{}^{2}}{(1+v^2)^2}}`$ $`2(\sqrt{1+v^2}+1)(\sqrt{1+v^2}1)=2v^2.`$ Moreover, $`2\sqrt{1+v^2}>(\sqrt{1+v^2}+1)`$, so one has $`r_{\mathrm{}}^2{\displaystyle \frac{v_{}^{}{}_{}{}^{2}}{(1+v^2)^{3/2}}}`$ $`v^2.`$ Taking square root and integrating, it yields $`r_{\mathrm{}}{\displaystyle _{r_0}^r_{\mathrm{}}}{\displaystyle \frac{v^{}}{(1+v^2)^{3/4}}}\mathrm{d}r`$ $`{\displaystyle _{r_0}^r_{\mathrm{}}}v\mathrm{d}r,`$ in which the left hand side is obviously convergent. The lower bound follows easily from the theorem. $`\mathrm{}`$ It can be seen that the area of negative part has a lower bound of order $`w_0^{}(r_{\mathrm{}}r_0)^2`$. We wish to establish that this dominates the area of the positive part. Let us consider the equation of $`v(r)=w(r)`$ for $`r>r_0`$ again, i.e., $`\left[{\displaystyle \frac{r^2v_{}^{}{}_{}{}^{2}}{(1+v^2)^{5/2}}}\right]^{}`$ $`=\left[{\displaystyle \frac{v^2}{(1+v^2)^{1/2}}}\right]^{}{\displaystyle \frac{2c_0v^2v^{}r}{1+v^2}}+{\displaystyle \frac{(c_0^2+\lambda )r^2vv^{}}{(1+v^2)^{1/2}}}+{\displaystyle \frac{pr^3v^{}}{2}}`$ $`\left[{\displaystyle \frac{v^2}{(1+v^2)^{1/2}}}\right]^{}+\left[{\displaystyle \frac{2|c_0|v^2r_{\mathrm{}}}{1+v^2}}+{\displaystyle \frac{(c_0^2+\lambda )r_{\mathrm{}}^2v}{(1+v^2)^{1/2}}}+{\displaystyle \frac{pr_{\mathrm{}}^3}{2}}\right]v^{}`$ $`\left[{\displaystyle \frac{v^2}{(1+v^2)^{1/2}}}\right]^{}+\left[2|c_0|r_{\mathrm{}}+(c_0^2+\lambda )r_{\mathrm{}}^2+{\displaystyle \frac{pr_{\mathrm{}}^3}{2}}\right]v^{}.`$ Therefore, after integrating from $`r_0`$, $`{\displaystyle \frac{r^2v_{}^{}{}_{}{}^{2}}{(1+v^2)^{5/2}}}`$ $`r_0^2v^{}(r_0)^2+{\displaystyle \frac{v^2}{(1+v^2)^{1/2}}}+\left[2|c_0|r_{\mathrm{}}+(c_0^2+\lambda )r_{\mathrm{}}^2+{\displaystyle \frac{pr_{\mathrm{}}^3}{2}}\right]v`$ $`r_0^2v^{}(r_0)^2+\left[1+2|c_0|r_{\mathrm{}}+(c_0^2+\lambda )r_{\mathrm{}}^2+{\displaystyle \frac{pr_{\mathrm{}}^3}{2}}\right]v`$ $`{\displaystyle \frac{r_0^2v_{}^{}{}_{}{}^{2}}{(1+v^2)^{5/2}}}`$ $`r_0^2v^{}(r_0)^2+\left[1+2|c_0|r_{\mathrm{}}+(c_0^2+\lambda )r_{\mathrm{}}^2+{\displaystyle \frac{p}{2}}r_{\mathrm{}}^3\right]v.`$ From this, we have proved a comparison between the growth of $`w^{}`$ and $`\left|w\right|`$. ###### Proposition 2.10. If $`w(r)`$ is a solution to equation (1) which blows down to $`\mathrm{}`$ at $`r_{\mathrm{}}`$, then $`{\displaystyle \frac{w_{}^{}{}_{}{}^{2}}{\left|w\right|(1+w^2)^{5/2}}}`$ is bounded. Continue with the analysis, after integration and applying Holder inequality, we have $`r_0^2`$ $`r_0^2\left({\displaystyle _{r_0}^r_{\mathrm{}}}{\displaystyle \frac{v^{}}{(1+v^2)^{5/4}}}1\right)^2r_0^2{\displaystyle _{r_0}^r_{\mathrm{}}}{\displaystyle \frac{v_{}^{}{}_{}{}^{2}}{(1+v^2)^{5/2}}}{\displaystyle _{r_0}^r_{\mathrm{}}}1`$ $`r_0^2v^{}(r_0)^2(r_{\mathrm{}}r_0)^2+(r_{\mathrm{}}r_0)\left[1+2|c_0|r_{\mathrm{}}+(c_0^2+\lambda )r_{\mathrm{}}^2+{\displaystyle \frac{p}{2}}r_{\mathrm{}}^3\right]{\displaystyle _{r_0}^r_{\mathrm{}}}v.`$ Letting $`x=r_{\mathrm{}}r_0`$ and using 2.6 that $`r_0^2<\frac{64}{\delta _+}w_0^{}`$ and $`v^{}(r_0)<2w_0^{}`$, we have (8) $$r_0^2\frac{64}{\delta _+}w_{0}^{}{}_{}{}^{3}x^2+x\left[a_0+a_1x+a_2x^2+\frac{p}{2}x^3\right]_{r_0}^r_{\mathrm{}}v,$$ where $`a_0`$ $`=1+2|c_0|r_0+(c_0^2+\lambda )r_0^2+{\displaystyle \frac{p}{2}}r_0^3,`$ $`a_1`$ $`=2|c_0|+2(c_0^2+\lambda )r_0+{\displaystyle \frac{3p}{2}}r_0^2\text{ and}`$ $`a_2`$ $`=(c_0^2+\lambda )+{\displaystyle \frac{3p}{2}}r_0.`$ We now claim that there exists $`C>0`$ independent of $`w_0^{}`$ such that $`_{r_0}^r_{\mathrm{}}vCw_0^{}`$ for sufficiently small $`w_0^{}`$. Otherwise, for any $`\epsilon >0`$, there is a $`w_0^{}<\epsilon `$ such that $$\epsilon w_0^{}>_{r_0}^r_{\mathrm{}}v.$$ Then by corollary 2.9, we have $$\epsilon w_0^{}>\frac{\delta }{2\sqrt{2}}r_0^2x^2.$$ Using the lower estimate of $`r_0^2`$ in terms of $`w_0^{}`$, we conclude that $$x^2<\epsilon \frac{2\sqrt{2}}{\delta }\frac{w_0^{}}{r_0^2}\frac{2\sqrt{2}\delta _+}{\delta }\epsilon .$$ Putting this into (8), we have $`c_1w_0^{}`$ $`c_2w_{0}^{}{}_{}{}^{3}\epsilon +c_3\epsilon ^{1/2}\{1+o(w_{0}^{}{}_{}{}^{1/2})+[2|c_0|+o(w_{0}^{}{}_{}{}^{1/2})]c_3\epsilon ^{1/2}+`$ $`+[c_0^2+\lambda +o(w_{0}^{}{}_{}{}^{1/2})]c_3^2\epsilon +{\displaystyle \frac{p}{2}}c_3^3\epsilon ^{3/2}\}\epsilon w_0^{},`$ for some positive constants $`c_1`$, $`c_2`$ and $`c_3`$. This is impossible since this implies $$0<c_1c_2w_{0}^{}{}_{}{}^{2}\epsilon +c_3\epsilon ^{3/2}\left[1+o(\epsilon ^{1/2})+o(w_{0}^{}{}_{}{}^{1/2})\right]0,$$ which is a contradiction. ###### Theorem 2.11. With the conditions of the theorems 2.3, 2.5, and 2.8, we have $$\mathrm{}<_0^r_{\mathrm{}}w(r)\mathrm{d}r<0$$ for $`w_0^{}`$ sufficiently small. ## 3. A Necessary Geometric Condition The fact that the surface has a reflection symmetry through the $`xy`$-plane actually provides some geometric conditions on the surface. Suppose we have a function $`z(r)`$ with $`w=z^{}=z_r`$ satisfying the equation (1), we must check that the surface obtained by revolving the function $`z(r)`$ is also stationary to the functional $``$. For this purpose, we may parametrize the surface by an even function $`r(z)`$ with $`z[a,a]`$, namely, $$\{\begin{array}{cc}\hfill 𝐗(\theta ,z)& =(r(z)\mathrm{cos}\theta ,r(z)\mathrm{sin}\theta ,z),(\theta ,z)[0,2\pi ]\times [a,a],\hfill \\ \hfill r(z)& =r(z),r(0)=r_{\mathrm{}}.\hfill \end{array}$$ In this section, we will use $`z_r`$ or $`w_r`$ to stand for derivatives wrt $`r`$, i.e., $`z^{}`$ and $`w^{}`$ previously in order to distinguish from derivatives wrt $`z`$ denoted by another subscript. The Helfrich functional for $`r(z)`$ and $`r(z)`$ near $`r_{\mathrm{}}`$ can be written as, $$\frac{}{2\pi }=\underset{\epsilon 0}{lim}_a^\epsilon +_\epsilon ^a\left(\left[(2H+c_0)^2\right]r\sqrt{1+r_z^2}+\frac{p}{2}r^2\right)\mathrm{d}z.$$ We then study the critical function of the functional. Assume that the variation $`\delta (r(z))=\phi (z)`$, then the above is written as $$\frac{\delta ()}{2\pi }=\underset{\epsilon 0}{lim}\left(F_2\phi _{zz}+F_1\phi _z+F_0\phi \right)\mathrm{d}z;$$ where (with similar calculations as those in the appendix) $$\begin{array}{cc}\hfill F_2& =\frac{2rr_{zz}}{(r_z^2+1)^{5/2}}\frac{2}{(r_z^2+1)^{3/2}}+\frac{2c_0r}{r_z^2+1};\hfill \\ \hfill F_1& =\frac{5rr_zr_{zz}^2}{(r_z^2+1)^{7/2}}\frac{r_z}{r(r_z^2+1)^{3/2}}+\frac{(c_0^2+\lambda )rr_z}{(r_z^2+1)^{1/2}}+\frac{6r_zr_{zz}}{(r_z^2+1)^{5/2}}\frac{4c_0rr_zr_{zz}}{(r_z^2+1)^2};\hfill \\ \hfill F_0& =\frac{r_{zz}^2}{(r_z^2+1)^{5/2}}\frac{1}{r^2(r_z^2+1)^{1/2}}+\frac{2c_0r_{zz}}{r_z^2+1}+(c_0^2+\lambda )\sqrt{r_z^2+1}+pr.\hfill \end{array}$$ Since $`F_2`$ and $`F_0`$ are even and $`F_1`$ is odd, after integrating by parts, we have $$\frac{\delta ()}{2\pi }=_a^a\left[(F_2)_{zz}(F_1)_z+F_0\right]\phi \mathrm{d}z\underset{\epsilon 0}{lim}\left[(F_2)_z(\epsilon )+F_1(\epsilon )\right]\left[\phi _z(\epsilon )\phi _z(\epsilon )\right].$$ The function $`r(z)`$ here is the inverse function of the function $`z(r)`$ in the previous sections. Thus, $`r_z`$ and $`r_{zz}`$ above can be written in terms of $`w`$ by chain rule. Since $`w(r)=z^{}(r)`$ is a solution to the equation (1), we obtain $$(F_2)_{zz}(F_1)_z+F_0=0.$$ Furthermore, we may denote $`\eta (r)\stackrel{\text{def}}{:==}(F_2)_z+F_1`$. Then, in terms of $`w`$, $$\eta =\frac{rw_r^2}{(1+w^2)^{5/2}}\frac{w^2}{r\sqrt{1+w^2}}2c_0w(c_0^2+\lambda )r\sqrt{1+w^2}+\frac{pr^2w}{2}.$$ Consequently, $`\delta ()=0`$ if and only if $`\eta (r)`$ is bounded in a neighborhood of $`r_{\mathrm{}}`$. Note that $$\eta _z=\frac{\eta _r}{w}=\frac{w_r^2}{w(1+w^2)^{5/2}}+\frac{w}{r^2\sqrt{1+w^2}}\frac{2c_0w_r}{w(1+w^2)}(c_0^2+\lambda )\frac{\sqrt{1+w^2}}{w}+pr$$ is bounded according to the analysis in the previous section, specifically by proposition 2.10. Therefore, $`\eta (r)`$ is also bounded and hence, we have established the following theorem. ###### Theorem 3.1. Let $`z=z(r)0`$ be a function on $`[0,r_{\mathrm{}}]`$ with $`z(r)=0`$ if and only if $`r=r_{\mathrm{}}`$. If $`w=z^{}`$ is a solution to the equation (1), then the surface obtained by revolving the curves $`z(r)`$ and $`z(r)`$ is stationary to the functional $``$. Furthermore, according to this boundary condition, by considering $`\underset{rr_{\mathrm{}}}{lim}{\displaystyle \frac{\eta (r)}{\sqrt{1+w^2}}}=0`$, there is a requirement on the geometry of the surface. ###### Proposition 3.2. Let $`\mathrm{\Sigma }`$ be an axisymmetric stationary surface of $``$. If $`r_{\mathrm{}}`$ is the radius of the circle of intersection of $`\mathrm{\Sigma }`$ and the reflection plane, then the Gaussian curvature of every point along this circle is given by $`K(r_{\mathrm{}})^2={\displaystyle \frac{1}{r_{\mathrm{}}}}Q\left({\displaystyle \frac{1}{r_{\mathrm{}}}}\right).`$ ## 4. Appendix: Derivation of the Main Equation Let the surface of revolution $`\mathrm{\Sigma }`$ be parametrized by $`𝐗(r,\theta )=(r\mathrm{cos}\theta ,r\mathrm{sin}\theta ,z(r))`$ for $`(r,\theta )[0,r_{\mathrm{}}]\times [0,2\pi ]`$. Then, with $`w(r)=z^{}(r)`$, the mean curvature $`H`$ is given by $$H=\frac{1}{2(1+w^2)}\left[w^{}+\frac{1}{r}w+\frac{1}{r}w^3\right].$$ The area element is $`2\pi r\sqrt{1+w^2}\mathrm{d}r`$. With the assumption that $`z(r_{\mathrm{}})=0`$, after integrating by parts, the volume enclosed by $`\mathrm{\Sigma }`$ is $`\pi {\displaystyle _0^r_{\mathrm{}}}r^2w(r)\mathrm{d}r.`$ Therefore, to find the critical point of $``$, we may consider the variation on $$\frac{1}{2\pi }=_0^r_{\mathrm{}}\left[(2H+c_0)^2+\lambda \right]r\sqrt{1+w^2}\mathrm{d}r\frac{p}{2}_0^r_{\mathrm{}}r^2w\mathrm{d}r.$$ Let $`\delta `$ be the variational operator and $`\delta (w)=\phi `$ where $`\phi (r)`$ is a smooth function with compact support. Then $`\delta /2\pi =_1+_2+_3`$ where $`_1`$ $`={\displaystyle _0^r_{\mathrm{}}}2(2H+c_0)\delta (H)r\sqrt{1+w^2}\mathrm{d}r;`$ $`_2`$ $`={\displaystyle _0^r_{\mathrm{}}}\left[(2H+c_0)^2+\lambda \right]r\delta (\sqrt{1+w^2})\mathrm{d}r`$ $`={\displaystyle _0^r_{\mathrm{}}}\left[(2H+c_0)^2+\lambda \right]r{\displaystyle \frac{rw\phi }{\sqrt{1+w^2}}}\mathrm{d}r;`$ $`_3`$ $`={\displaystyle \frac{p}{2}}{\displaystyle _0^r_{\mathrm{}}}r^2\delta (w)\mathrm{d}r={\displaystyle \frac{p}{2}}{\displaystyle _0^r_{\mathrm{}}}r^2\phi \mathrm{d}r.`$ We also have $`\delta (2H)`$ $`={\displaystyle \frac{1}{(1+w^2)^{5/2}}}\left[(1+w^2)\phi ^{}+{\displaystyle \frac{1}{r}}(1+w^2)\phi 3ww^{}\phi \right];`$ $`(2H)^{}`$ $`={\displaystyle \frac{1}{(1+w^2)^{5/2}}}\left[(1+w^2)w^{\prime \prime }3ww_{}^{}{}_{}{}^{2}+{\displaystyle \frac{1}{r}}(1+w^2)w^{}{\displaystyle \frac{1}{r^2}}w(1+w^2)^2\right].`$ Therefore, $`_1`$ $`={\displaystyle _0^r_{\mathrm{}}}{\displaystyle \frac{2r(2H+c_0)}{(1+w^2)^2}}\left[(1+w^2)\phi ^{}+{\displaystyle \frac{1}{r}}(1+w^2)\phi 3ww^{}\phi \right]`$ $`={\displaystyle _0^r_{\mathrm{}}}{\displaystyle \frac{}{r}}\left[{\displaystyle \frac{2r(2H+c_0)}{(1+w^2)}}\right]\phi \mathrm{d}r+{\displaystyle _0^r_{\mathrm{}}}{\displaystyle \frac{2r(2H+c_0)}{(1+w^2)}}\phi \mathrm{d}r`$ $`{\displaystyle _0^r_{\mathrm{}}}{\displaystyle \frac{6rww^{}(2H+c_0)}{(1+w^2)^2}}\phi \mathrm{d}r`$ $`={\displaystyle _0^r_{\mathrm{}}}\left[{\displaystyle \frac{2r(2H)^{}}{(1+w^2)}}{\displaystyle \frac{2rww^{}(2H+c_0)}{(1+w^2)^2}}\right]\phi \mathrm{d}r`$ $`={\displaystyle _0^r_{\mathrm{}}}[{\displaystyle \frac{2rw^{\prime \prime }}{(1+w^2)^{5/2}}}+{\displaystyle \frac{6rww_{}^{}{}_{}{}^{2}}{(1+w^2)^{7/2}}}{\displaystyle \frac{2w^{}}{(1+w^2)^{5/2}}}+{\displaystyle \frac{2w}{r(1+w^2)^{3/2}}}`$ $`{\displaystyle \frac{2c_0rww^{}}{(1+w^2)^2}}{\displaystyle \frac{2rww_{}^{}{}_{}{}^{2}}{(1+w^2)^{7/2}}}{\displaystyle \frac{2w^2w^{}}{(1+w^2)^{5/2}}}]\phi \mathrm{d}r.`$ The quantity $`_2`$ is expanded into the following. $`_2`$ $`={\displaystyle _0^r_{\mathrm{}}}{\displaystyle \frac{\left[(2H+c_0)^2+\lambda \right]rw}{(1+w^2)^{1/2}}}\phi \mathrm{d}r`$ $`={\displaystyle _0^r_{\mathrm{}}}[{\displaystyle \frac{rw}{(1+w^2)^{7/2}}}(w^{}+{\displaystyle \frac{1}{r}}w(1+w^2))^2`$ $`+{\displaystyle \frac{2c_0rw}{(1+w^2)^2}}(w^{}+{\displaystyle \frac{1}{r}}w(1+w^2))+{\displaystyle \frac{(c_0^2+\lambda )rw}{\sqrt{1+w^2}}}]\phi \mathrm{d}r`$ $`={\displaystyle _0^r_{\mathrm{}}}[{\displaystyle \frac{rww_{}^{}{}_{}{}^{2}}{(1+w^2)^{7/2}}}+{\displaystyle \frac{2w^2w^{}}{(1+w^2)^{5/2}}}+{\displaystyle \frac{w^3}{r(1+w^2)^{3/2}}}`$ $`+{\displaystyle \frac{2c_0rww^{}}{(1+w^2)^2}}+{\displaystyle \frac{2c_0w^2}{(1+w^2)}}+{\displaystyle \frac{(c_0^2+\lambda )rw}{\sqrt{1+w^2}}}]\phi \mathrm{d}r.`$ Summing the above quantities, we have $`{\displaystyle \frac{1}{2\pi }}\delta ()`$ $`=_1+_2+_3`$ $`={\displaystyle _0^r_{\mathrm{}}}[{\displaystyle \frac{2rw^{\prime \prime }}{(1+w^2)^{5/2}}}+{\displaystyle \frac{5rww_{}^{}{}_{}{}^{2}}{(1+w^2)^{7/2}}}{\displaystyle \frac{2w^{}}{(1+w^2)^{5/2}}}+{\displaystyle \frac{2w+w^3}{r(1+w^2)^{3/2}}}`$ $`+{\displaystyle \frac{2c_0w^2}{(1+w^2)}}+{\displaystyle \frac{(c_0^2+\lambda )rw}{(1+w^2)^{1/2}}}{\displaystyle \frac{pr^2}{2}}]\phi \mathrm{d}r.`$ Therefore, the variational equation is given by $`(\text{integrand})+\text{constant}=0`$. With the assumption that $`w(0)=0`$, taking $`r0`$, one gets $$2w^{}(0)+\underset{r0}{lim}\frac{2w}{r}+\text{constant}=0.$$ This concludes the derivation of the variational equation.
warning/0001/math0001122.html
ar5iv
text
# Convergence of Bieberbach polynomials in domains with interior cusps ## 1. Introduction Let $`G`$ be a bounded Jordan domain, $`z_0G`$. Define the Bergman space $`L_2(G)`$ as the space of square integrable analytic functions with norm $$f_2:=\left(_G|f(z)|^2𝑑x𝑑y\right)^{1/2}.$$ We also use the uniform norm on $`\overline{G}`$ in the sequel: $$f_{\mathrm{}}:=\underset{zG}{sup}|f(z)|.$$ The Bieberbach polynomial $`B_n(z),\mathrm{deg}B_nn`$, is the solution of the following extremal problem : (1.1) $$B_n^{}_2=\underset{P_n𝒫_n()}{inf}\{P_n^{}_2:P_n(z_0)=0,P_n^{}(z_0)=1\},$$ where $`𝒫_n()`$ is the class of algebraic polynomials of degree at most $`n`$, with complex coefficients. The conformal mapping $`\phi :GD_{R_0}:=\{z:|z|<R_0\}`$, normalized by $`\phi (z_0)=0`$ and $`\phi ^{}(z_0)=1`$, solves the same extremal problem in the class of all analytic functions $`f`$ in $`G`$, satisfying $`f(z_0)=0`$ and $`f^{}(z_0)=1`$. Here, $`R_0`$ is the inner conformal radius of the domain $`G`$ with respect to $`z_0.`$ Moreover, (1.2) $$\phi ^{}B_n^{}_2=\underset{P_n𝒫_n()}{inf}\{\phi ^{}P_n^{}_2:P_n(z_0)=0,P_n^{}(z_0)=1\}$$ (see, e.g., or ). It is clear from (1.2) that (1.3) $$\underset{n\mathrm{}}{lim}B_n^{}(z)=\phi ^{}(z)\text{ and }\underset{n\mathrm{}}{lim}B_n(z)=\phi (z),zG,$$ because polynomials are dense in $`L_2(G)`$. A more delicate fact of the uniform convergence of $`B_n(z)`$ to $`\phi (z)`$ on $`\overline{G}`$ was first observed by Keldysh in 1939 , for the domains $`G`$ with sufficiently smooth boundaries. He also constructed an example of a starlike domain, bounded by a piecewise analytic curve with one singular point, where Bieberbach polynomials diverge. A considerable progress in the area has been achieved by Mergelyan , Suetin , Simonenko , Andrievskii - and Gaier -. In particular, Andrievskii proved that the uniform convergence of Bieberbach polynomials holds on $`\overline{G}`$, where $`G`$ is any quasidisk, and Gaier - showed that the rate of this uniform convergence is quite close to the best possible rate in uniform polynomial approximation of the conformal mapping $`\phi `$. It is well known that a quasiconformal curve does not allow zero angles (cusps). The first results on the uniform convergence of Bieberbach polynomials in domains with cusps were obtained by Andrievskii -. Pritsker developed his approach to improve these results for domains with certain interior zero angles (outward pointing cusps). The interior zero angles seem to play a special role in this problem, as Keldysh’s counterexample, although being implicit, but gives impression that his piecewise analytic boundary curve has an outward cusp as its only singular point (see ). We confirm this by constructing an example on the divergence of Bieberbach polynomials at an outward pointing cusp (see Theorem 2.2 and its proof). An interesting problem arising here is to find the critical order of tangency at this interior zero angle, separating the convergent behavior of Bieberbach polynomials, exhibited below in Theorem 2.1, from the divergent one for sufficiently thin cusps. This would give a rather complete answer to the old question on the geometry of domains with uniform convergence of Bieberbach polynomials. Introducing the area orthonormal polynomials $`\{K_n(z)\}_{n=0}^{\mathrm{}}`$, such that (1.4) $$_GK_m(z)\overline{K_n(z)}𝑑x𝑑y=\{\begin{array}{c}1,m=n,\hfill \\ 0,mn,\hfill \end{array}$$ one can find the following representation for Bieberbach polynomials \[10, p. 34\]: (1.5) $$B_n(z)=\frac{{\displaystyle \underset{k=0}{\overset{n1}{}}}\overline{K_k(z_0)}{\displaystyle _{z_0}^z}K_k(t)𝑑t}{{\displaystyle \underset{k=0}{\overset{n1}{}}}|K_k(z_0)|^2},n.$$ This gives a constructive method for generating Bieberbach polynomials via the Gram-Schmidt orthonormalization process and for numerical approximation of the conformal mapping $`\phi `$ (see ). In addition, (1.5) indicates the connection with the Bergman kernel function (1.6) $$K(z,z_0)=\underset{k=0}{\overset{\mathrm{}}{}}\overline{K_k(z_0)}K_k(z)=\frac{\phi ^{}(z)}{\pi R_0^2},z,z_0G.$$ It is clear from (1.6) and $`\phi ^{}(z_0)=1`$ that (1.7) $$K(z_0,z_0)=\underset{k=0}{\overset{\mathrm{}}{}}|K_k(z_0)|^2=\frac{1}{\pi R_0^2}.$$ Thus, many problems on the convergence of Bieberbach polynomials are equivalent to those on the convergence of the integrated bilinear series of (1.6). ## 2. Convergence and Divergence Results Let $`\tau `$ be a conformal mapping of the unit disk $`D`$ onto a quasidisk (cf. ). We say that a Jordan arc $`\gamma `$ is quasianalytic if $`\gamma =\tau ([1,1])`$ for such a mapping $`\tau `$ (quasianalytic arcs were introduced in ). It is known that a quasianalytic arc is rectifiable and quasismooth, i.e., it satisfies the following chord-arc condition, by Lavrentiev: $$|\gamma (z_1,z_2)|M|z_1z_2|,z_1,z_2\gamma ,$$ where $`|\gamma (z_1,z_2)|`$ is the length of a subarc $`\gamma (z_1,z_2)\gamma ,`$ with the endpoints $`z_1,z_2,`$ and $`M1`$ is a constant, depending only on $`\gamma .`$ A Jordan curve is said to be piecewise quasianalytic, if it consists of a finite number of quasianalytic arcs. Suppose that $`G`$ is bounded by a piecewise quasianalytic curve $`L=G`$, with the quasianalytic arcs joining at the points $`\{z_j\}_{j=1}^mL`$. Two quasianalytic arcs $`L_jL`$ and $`L_{j+1}L`$, meeting at $`z_j`$, form an $`x^p`$-type interior zero angle, if there exists a neighborhood of $`z_j`$ such that in a local coordinate system, with the origin at $`z_j`$, we have $$(x,y)L_jc_1x^Pyc_2x^p$$ and $$(x,y)L_{j+1}c_2x^pyc_1x^P,$$ where $`Pp>1`$ and $`c_1,c_2>0.`$ With these notations, our result on the convergence of Bieberbach polynomials is stated below. ###### Theorem 2.1. If $`G`$ is piecewise quasianalytic, with $`x^p`$-type interior zero angles at the joint points, then there exist $`q=q(G),r=r(G),0<q,r<1,`$ and $`C=C(G)>0`$ such that (2.1) $$\phi B_n_{\mathrm{}}Cq^{n^r},n.$$ It is worth noting that one cannot have $`r=1`$ in Theorem 2.1, as this would imply that $`\phi `$ is analytic on $`\overline{G}`$, by a well known result (see, e.g., \[10, p. 27\]), which is obviously not the case (cf. Theorem 3.1 below). A companion divergence result is based on Keldysh’s construction in , but its geometry is made more explicit here. ###### Theorem 2.2. There exists a domain with piecewise smooth boundary and one outward pointing cusp, such that Bieberbach polynomials diverge at this cusp. Furthermore, the boundary of this domain is analytic outside of any neighborhood of the cusp point. One might speculate that the interior zero angle in Theorem 2.2 has an exponential order of tangency at the cusp point. It would be very interesting to find the critical order of tangency at this interior zero angle, separating the convergent behavior of Bieberbach polynomials in Theorem 2.1, from the divergent one in Theorem 2.2. ###### Remark 2.3. It is also interesting to note that the boundary of the domain in Theorem 2.2 (or in Keldysh’s counterexample) cannot be piecewise analytic, as is claimed in . In other words, the boundary arc, with endpoints meeting at the only irregular boundary point, cannot be an analytic arc, which is the image of a segment under a mapping, analytic in a domain containing this segment inside. Indeed, if this arc is analytic then we can define the angle at the irregular boundary point, by using one-sided tangents. In the case this angle is non-zero, we see that the boundary of our domain is quasiconformal, so that the associated Bieberbach polynomials must converge uniformly by . Thus divergence is only possible if we have a zero angle at the irregular point, which translates into an outward pointing cusp in Keldysh’s construction. However, an analytic arc can only form an $`x^p`$-type zero angle, because it cannot have an arbitrarily high order of contact, as we show in Section 4. Hence we get the uniform convergence of Bieberbach polynomials again, by Theorem 2.1! ## 3. Continuity and Differentiability of a Conformal Mapping at a Cusp The results on the behavior of $`\phi `$ at an interior zero angle, obtained in this paper, may be of independent interest. We summarize them below for convenience of the reader. ###### Theorem 3.1. Suppose that $`G`$ has an $`x^p`$-type interior zero angle at $`zG`$, which is formed by two quasianalytic arcs. Then there exist constants $`C,c>0`$ such that (3.1) $$|\phi (t)\phi (z)|C\mathrm{exp}\left(\frac{c}{|tz|^{p1}}\right),tz,t\overline{G}.$$ Furthermore, (3.2) $$\underset{\genfrac{}{}{0pt}{}{tz}{t\overline{G}\{z\}}}{lim}\frac{\phi ^{(k)}(t)}{|tz|^m}=0,k,m.$$ ###### Corollary 3.2. Suppose that $`G`$ is piecewise quasianalytic, with $`x^p`$-type interior zero angles at the joint points. Then $`\phi C^{\mathrm{}}(\overline{G})`$, i.e., (3.3) $$\phi ^{(k)}C(\overline{G}),k.$$ ## 4. Proofs Let $`C,c,c_1,c_2,\mathrm{}`$ denote positive constants, not necessarily the same at different places. Writing $`ab`$, we mean that $`ac_1b`$ for a constant $`c_1`$, which doesn’t depend on $`a`$ and $`b`$. The relation $`ab`$ indicates that $`c_2bac_1b`$, where $`c_1,c_2`$ are independent of $`a`$ and $`b`$. Our proofs heavily rely on the distortion properties of conformal and quasiconformal mappings, where we start with the following lemma (see Andrievskii \[6, pp. 97-98\]). ###### Lemma 4.1. Let $`w=F(\zeta )`$ be a $`K`$-quasiconformal mapping of the plane onto itself, such that $`F(\mathrm{})=\mathrm{}`$, $`\zeta _j`$, $`w_j=F(\zeta _j)`$ $`(j=1,2,3)`$, and $`|w_1w_2|c_1|w_1w_3|`$. Then $`|\zeta _1\zeta _2|c_2|\zeta _1\zeta _3|`$ and (4.1) $$c_3\left|\frac{w_1w_3}{w_1w_2}\right|^{1/K}\left|\frac{\zeta _1\zeta _3}{\zeta _1\zeta _2}\right|c_4\left|\frac{w_1w_3}{w_1w_2}\right|^K,$$ where $`c_j=c_j(c_1,K)`$, $`j=2,3,4`$. We next introduce the arc $`l_{q,a}D`$, with the endpoints at $`\pm 1`$, as the union of two arcs $`\{w=x+iy:y=(x+1)^q/a,1<x0\}`$ and $`\{w=x+iy:y=(1x)^q/a,0x<1\},q,a>1.`$ If $`\tau `$ is the conformal map of $`D`$, defining a quasianalytic arc of the boundary $`L_1G`$ with the cusp points $`z_1=\tau (1)`$ and $`z_2=\tau (1)`$, then we set $$L_{q,a}:=\tau (l_{q,a}).$$ We assume here that $`\tau `$ is extended to a $`K`$-quasiconformal homeomorphism of the complex plane onto itself, with infinity as a fixed point. It is clear that $`L_{q,a}`$ is an arc connecting $`z_1`$ and $`z_2`$ in the exterior of $`G`$. We give estimates for the distance $`d(\zeta ,L_1)`$ from $`\zeta L_{q,a}`$ to $`L_1`$, $`q,a>1`$, and also for the distance $`d(\zeta ,L_{q,a})`$ from $`\zeta L_1`$ to $`L_{q,a}`$. ###### Lemma 4.2. If $`\zeta L_{q,a}`$ then (4.2) $$\underset{j=1,2}{\mathrm{min}}|\zeta z_j|^{(q1)K^2+1}d(\zeta ,L_1)\underset{j=1,2}{\mathrm{min}}|\zeta z_j|^{(q1)/K^2+1}.$$ Similarly, if $`\zeta L_1`$ then (4.3) $$\underset{j=1,2}{\mathrm{min}}|\zeta z_j|^{(q1)K^2+1}d(\zeta ,L_{q,a})\underset{j=1,2}{\mathrm{min}}|\zeta z_j|^{(q1)/K^2+1}.$$ ###### Proof. Set $`t=\tau ^1(\zeta )`$ and $`x=\mathrm{}t`$. Assume that $`x0`$. Since $`|tx|<|t1|`$, we obtain that $$|t1|^{1q}\left|\frac{t1}{tx}\right|\left|\frac{\zeta z_2}{\zeta \tau (x)}\right|^K\left(\frac{|\zeta z_2|}{d(\zeta ,L)}\right)^K,$$ by Lemma 4.1. Therefore $$d(\zeta ,L)|t1|^{(q1)/K}|\zeta z_2|.$$ Applying (4.1) again, with $`\zeta _1=z_2,\zeta _2=\zeta \text{ and }\zeta _3=z_1`$, we have (4.4) $$|\zeta z_2|^K|t1||\zeta z_2|^{1/K},$$ which gives the right hand side of (4.2), by the previous inequality. Let $`\zeta ^{}L`$ be such that $`d(\zeta ,L)=|\zeta \zeta ^{}|`$, and set $`t^{}=\tau ^1(\zeta ^{})`$. Assume that $`\mathrm{}t^{}0`$. Since $`|\zeta \zeta ^{}||\zeta z_2|`$, Lemma 4.1 yields $$\frac{|\zeta z_2|}{d(\zeta ,L)}=\left|\frac{\zeta z_2}{\zeta \zeta ^{}}\right|\left|\frac{t1}{tt^{}}\right|^K\left|\frac{t1}{tx}\right|^K|t1|^{(1q)K}.$$ Combining the above estimate with (4.4), we also prove the left hand side of (4.2). The estimates in (4.3) are obtained by an analogous argument. ∎ ###### Remark 4.3. If we consider $`\overline{l}_{q,a}:=\{\overline{z}:zl_{q,a}\}D`$ and set $`\overline{L}_{q,a}:=\tau (\overline{l}_{q,a}),`$ then (4.2) holds for any $`\zeta \overline{L}_{q,a}`$ as well. We now construct an analytic extension of the conformal mapping $`\phi `$ into a domain $`\stackrel{~}{G}`$ containing $`G`$. ###### Lemma 4.4. Suppose that $`G`$ is piecewise quasianalytic, with $`x^p`$-type interior zero angles at the joint points. Then the mapping $`\phi `$ can be continued conformally into a domain $`\stackrel{~}{G}`$, with the rectifiable boundary $`\stackrel{~}{G}`$ that consists of quasismooth arcs $`L_{q,a}`$, connecting the cusp points $`\{z_j\}_{j=1}^m`$, such that $`G\stackrel{~}{G}`$ and $`G\stackrel{~}{G}=\{z_j\}_{j=1}^m`$. Furthermore, there exist constants $`C,c>0`$ such that (4.5) $$|\phi (z)\phi (z_j)|C\mathrm{exp}\left(\frac{c}{|zz_j|^{p1}}\right),z\stackrel{~}{G},$$ where $`j=1,\mathrm{},m.`$ ###### Proof. Let $`L_j`$, $`j=1,2,\mathrm{},m,`$ be the quasianalytic arc of $`L=G`$, connecting $`z_j`$ and $`z_{j+1}`$, which is defined by the corresponding conformal map $`\tau _j`$. Denote the domain, bounded by the arcs $`l_{q,a}`$ and $`\overline{l}_{q,a}`$, by $`S_{q,a}`$. It follows from (4.2) and Remark 4.3 that $$\tau _j(S_{q,a})\tau _k(S_{q,a})=\mathrm{}\text{ for }jk,$$ provided we choose $`a>0`$ to be sufficiently small and $`q`$ to satisfy $`(q1)/K^2+1>P`$. On defining $$\stackrel{~}{G}:=G\left[\underset{j=1}{\overset{m}{}}\tau _j(S_{q,a})\right],$$ we extend the conformal mapping $`\phi `$ into $`\stackrel{~}{G}`$ using the standard reflection principle: (4.6) $$\phi (z):=\frac{R_0^2}{\overline{\phi \left[\tau _j\left(\overline{\tau _j^1(z)}\right)\right]}},z\tau _j(S_{q,a})\backslash \overline{G},$$ where $`j=1,\mathrm{},m.`$ We next proceed to proving (4.5), where we use the method of moduli of curve families (the method of extremal length). There is no loss of generality in assuming that $`z_j=0G`$ and that $`GD_R(0)W`$ for some $`R>0`$, where the wedge $`W`$ is defined by $`W:=\{w=x+iy:|y|<c_1x^p,x>0\},`$ with $`p>1.`$ Fix a point $`aG`$ and set $`d=\mathrm{min}(|a|/2,R).`$ Consider a point $`z\overline{G},`$ such that $`|z|<d`$, and a family of curves $`\mathrm{\Gamma }`$, separating points $`0`$ and $`z`$ from the point $`a`$ in $`G`$. We need to estimate the module of $`\mathrm{\Gamma }`$, denoted by $`m(\mathrm{\Gamma })`$, from below. This is accomplished with the help of an auxiliary family of curves $`\mathrm{\Gamma }^{}`$, which consists of the circular arcs $`\gamma (r):=\{|w|=r\}W`$, with the radius $`r`$ varying from $`c_2|z|`$ to $`d`$, $`c_2|z|<r<d.`$ Since the boundary arcs $`L_j`$ and $`L_{j+1}`$, meeting at $`z_j=0`$, are quasismooth, we can choose $`c_2`$ such that each curve from $`\mathrm{\Gamma }^{}`$ contains a curve from $`\mathrm{\Gamma }.`$ It follows by the comparison principle (see Theorem 4-1 in \[1, p. 54\]) that (4.7) $$m(\mathrm{\Gamma })m(\mathrm{\Gamma }^{}).$$ Let $`\theta (r)`$ be the angular measure of the arc $`\gamma (r)\mathrm{\Gamma }^{}.`$ It is known (cf. Theorem 2.6 in \[18, p. 77\]) that (4.8) $$m(\mathrm{\Gamma }^{})=_{c_2|z|}^d\frac{dr}{r\theta (r)}.$$ Using a simple estimate $`\theta (r)2c_1r^{p1},p>1`$, we conclude by (4.7) and (4.8) that $$m(\mathrm{\Gamma })_{c_2|z|}^d\frac{dr}{2c_1r^p}=\frac{1}{2(p1)c_1}\left(\frac{1}{(c_2|z|)^{p1}}\frac{1}{d^{p1}}\right).$$ Hence (4.9) $$|\phi (z)\phi (0)|e^{\pi m(\mathrm{\Gamma })}\mathrm{exp}\left(\frac{c}{|z|^{p1}}\right),z\overline{G},$$ by Theorem 1 of \[7, p. 290\] (see also \[6, p. 34\]). In the case $`z\stackrel{~}{G}\overline{G}`$, we obtain from (4.6) and (4.9) that $$|\phi (z)\phi (0)|\mathrm{exp}\left(\frac{c}{\left|\tau _j\left(\overline{\tau _j^1(z)}\right)\right|^{p1}}\right),$$ with a different $`c`$. Applying (4.1) with $`\zeta _1=0,\zeta _2=\tau _j\left(\overline{\tau _j^1(z)}\right),\zeta _3=z`$ and $`w_1=1,w_2=\overline{\tau _j^1(z)},w_3=\tau _j^1(z),`$ we conclude that $$\left|\tau _j\left(\overline{\tau _j^1(z)}\right)\right||z|,$$ which implies (4.5) by the previous inequality. ∎ ###### Lemma 4.5. Let $`G`$ be a Jordan domain, which is symmetric in the real axis, and let $`z_0=0G`$. Assume that $`\xi G`$ is real and $`G\{z:\mathrm{}z<\xi \}`$. If the conformal mapping $`\phi `$ is not analytic on $`\overline{G}`$, then (4.10) $$\underset{n\mathrm{}}{lim\; sup}|B_n(x)|=\mathrm{},x>\xi .$$ Proof. Suppose to the contrary of (4.10) that there exists $`x_0>\xi `$ such that the sequence $`\{B_n(x_0)\}_{n=0}^{\mathrm{}}`$ is bounded. Then we obtain from (1.5) and (1.7) that the following sequence is also bounded: $$\underset{k=0}{\overset{n1}{}}\overline{K_k(0)}_0^{x_0}K_k(t)𝑑t,n.$$ This implies, in turn, that (4.11) $$\left|\overline{K_n(0)}_0^{x_0}K_n(t)𝑑t\right|<C,n,$$ for some constant $`C>0`$. Note that the orthonormal polynomials $`K_n(z)`$ have real coefficients for any $`n`$, because $`G`$ is symmetric about the real axis. Furthermore, we follow the usual convention in Gram-Schmidt orthonormalization that the leading coefficient of $`K_n(z)`$ is positive for any $`n`$. It follows that each $`K_n(x)`$ is real valued for real $`x`$, and is positive for $`x+\mathrm{}`$. Since the zeros of $`K_n(z)`$ are contained in the convex hull of $`\overline{G}`$ (see \[24, p. 31\]), we conclude that $`K_n(x)`$ has no zeros for $`x>\xi `$ and that (4.12) $$K_n(x)>0,x>\xi ,n.$$ Using Theorem 1.1.4 of \[24, p. 4\], we obtain that (4.13) $$\underset{n\mathrm{}}{lim\; sup}K_n_{\mathrm{}}^{1/n}1$$ and (4.14) $$\underset{n\mathrm{}}{lim}|K_n(x)|^{1/n}=e^{g_\mathrm{\Omega }(x,\mathrm{})}>1,x>\xi ,$$ where $`g_\mathrm{\Omega }(x,\mathrm{})`$ is the Green function of $`\mathrm{\Omega }:=\overline{}\backslash \overline{G}`$ with pole at $`\mathrm{}`$. Combining (4.12)-(4.14) gives that $$\underset{n\mathrm{}}{lim\; inf}\left|_0^{x_0}K_n(t)𝑑t\right|^{1/n}>1$$ and that (4.15) $$\underset{n\mathrm{}}{lim\; sup}|K_n(0)|^{1/n}<1,$$ by (4.11). Hence the conformal mapping $`\phi `$ must have an analytic continuation through $`G`$, by (4.15) and Theorem 2.1 of , which contradicts our assumption. ∎ Proof of Theorem 2.1. We use a known method, based on the extremal property of Bieberbach polynomials (1.2). Namely, we first estimate the quantity $`\phi ^{}B_n^{}_2`$, and then proceed to the uniform norm case, to prove (2.1). Recall that the conformal mapping $`\phi `$ can be continued, by Lemma 4.4, into a larger domain $`\stackrel{~}{G}`$, whose boundary consists of quasismooth arcs connecting the cusp points $`\{z_j\}_{j=1}^mG`$. Let $`\gamma _j`$ be a subarc of $`\stackrel{~}{G}`$, with the endpoints $`z_j`$ and $`z_{j+1}`$, and let $`\zeta _j\gamma _j`$ be a fixed point, $`j=1,\mathrm{},m.`$ Note that $`\zeta _j`$ divides $`\gamma _j`$ into $`\gamma _j^1`$ and $`\gamma _j^2`$, so that $`\stackrel{~}{G}=_{j=1}^m_{i=1}^2\gamma _j^i`$. Since $`\stackrel{~}{G}`$ is rectifiable, we have by Cauchy’s integral formula that $`\phi (z)`$ $`=`$ $`{\displaystyle \frac{1}{2\pi i}}{\displaystyle _{\stackrel{~}{G}}}{\displaystyle \frac{\phi (\zeta )}{\zeta z}}𝑑\zeta ={\displaystyle \frac{1}{2\pi i}}{\displaystyle \underset{j=1}{\overset{m}{}}}{\displaystyle \underset{i=1}{\overset{2}{}}}{\displaystyle _{\gamma _j^i}}{\displaystyle \frac{\phi (\zeta )}{\zeta z}}𝑑\zeta `$ $`=`$ $`{\displaystyle \frac{1}{2\pi i}}{\displaystyle \underset{j=1}{\overset{m}{}}}{\displaystyle \underset{i=1}{\overset{2}{}}}{\displaystyle _{\gamma _j^i}}{\displaystyle \frac{\phi (\zeta )\phi (z_{j+i1})}{\zeta z}}𝑑\zeta +{\displaystyle \frac{1}{2\pi i}}{\displaystyle \underset{j=1}{\overset{m}{}}}\phi (z_j)\mathrm{log}{\displaystyle \frac{\zeta _{j1}z}{\zeta _jz}},`$ for any $`zG`$, where we assume that $`\zeta _0=\zeta _m`$. It follows that $`\phi ^{}(z)`$ $`=`$ $`{\displaystyle \frac{1}{2\pi i}}{\displaystyle \underset{j=1}{\overset{m}{}}}{\displaystyle \underset{i=1}{\overset{2}{}}}{\displaystyle _{\gamma _j^i}}{\displaystyle \frac{\phi (\zeta )\phi (z_{j+i1})}{(\zeta z)^2}}𝑑\zeta `$ $`+`$ $`{\displaystyle \frac{1}{2\pi i}}{\displaystyle \underset{j=1}{\overset{m}{}}}\phi (z_j)\left({\displaystyle \frac{1}{\zeta _jz}}{\displaystyle \frac{1}{\zeta _{j1}z}}\right),zG.`$ Observe that the second sum $$f(z):=\frac{1}{2\pi i}\underset{j=1}{\overset{m}{}}\phi (z_j)\left(\frac{1}{\zeta _jz}\frac{1}{\zeta _{j1}z}\right)$$ represents a function, analytic on $`\overline{G}.`$ Hence there exists a sequence of polynomials $`\{p_n\}_{n=1}^{\mathrm{}}`$ and a number $`R>1`$ such that (4.17) $$fp_n_{\mathrm{}}CR^n,n,$$ (see, e.g., Theorem 4 in \[10, p. 27\]). Consequently, our problem of approximating $`\phi ^{}(z)`$ by polynomials reduces to approximating functions of the following form $$g_{i,j}(z):=_{\gamma _j^i}\frac{\phi (\zeta )\phi (z_{j+i1})}{(\zeta z)^2}𝑑\zeta ,$$ in view of (4). Furthermore, we can consider approximation in the uniform norm and then pass to $`L_2(G)`$ norm, which suffices for our purposes. We now set (4.18) $$g(z):=g_{1,1}(z)=_\gamma \frac{\phi (\zeta )\phi (z_1)}{(\zeta z)^2}𝑑\zeta ,\gamma :=\gamma _1^1,$$ and study the approximation of this function only, as the other functions $`g_{i,j}(z)`$ are handled similarly. Let $`\mathrm{\Phi }:\mathrm{\Omega }D^{}`$ be a conformal map of $`\mathrm{\Omega }:=\overline{}\overline{G}`$ onto $`D^{}:=\{w:|w|>1\}`$, satisfying the conditions $`\mathrm{\Phi }(\mathrm{})=\mathrm{}`$ and $`\mathrm{\Phi }^{}(\mathrm{})>0.`$ Define the level curves of $`\mathrm{\Phi }`$ by $$L_R:=\{z\overline{\mathrm{\Omega }}:|\mathrm{\Phi }(z)|=R\},R1,$$ where we set $`L:=L_1=\mathrm{\Omega }=G.`$ Let $`G_R:=IntL_R,R>1,`$ be the domain bounded by $`L_R.`$ Clearly, if $`R`$ is sufficiently close to 1, then $`\gamma L_R\mathrm{}.`$ Denote $`\gamma ^{}:=\gamma \overline{G_R}`$ and $`\gamma ^{\prime \prime }:=\gamma \gamma ^{}`$, so that $`\gamma ^{\prime \prime }`$ lies exterior to $`L_R`$. Hence the function (4.19) $$h_2(z):=_{\gamma ^{\prime \prime }}\frac{\phi (\zeta )\phi (z_1)}{(\zeta z)^2}𝑑\zeta $$ is holomorphic in $`G_R`$ and is well approximable by polynomials. Namely, we obtain from Theorem 3 of \[23, p. 145\] that there exists a sequence of polynomials $`\{p_n\}_{n=1}^{\mathrm{}}`$ such that (4.20) $$h_2p_n_{\mathrm{}}C\frac{n}{(r1)^2}\underset{zG_r}{\mathrm{max}}|h_2(z)|r^n,n,$$ where $`C`$ is an absolute constant and $`r<R`$. On choosing $`R=1+2n^s`$ and $`r=1+n^s`$, with $`s(0,1)`$, we estimate $`\underset{zG_r}{\mathrm{max}}|h_2(z)|`$ $``$ $`{\displaystyle _{\gamma ^{\prime \prime }}}{\displaystyle \frac{|\phi (\zeta )\phi (z_1)|}{|\zeta z|^2}}|d\zeta |{\displaystyle \frac{1}{\underset{zG_r,\zeta \gamma ^{\prime \prime }}{\mathrm{min}}|\zeta z|^2}}`$ $``$ $`{\displaystyle \frac{1}{[d(L_R,L_r)]^2}},`$ where $`d(L_R,L_r)`$ is the distance between $`L_R`$ and $`L_r.`$ Note that $$d(L_R,L_r)c(Rr)^2,$$ by a result of Loewner (see \[6, p. 61\]), which implies $$d(L_R,L_r)n^{2s}.$$ We conclude that $$\underset{zG_r}{\mathrm{max}}|h_2(z)|n^{4s},$$ and, using (4.20), (4.21) $$h_2p_n_{\mathrm{}}n^{1+2s+4s}(1+n^s)^nn^{1+6s}e^{n^{1s}},n.$$ Introducing a companion function $$h_1(z):=_\gamma ^{}\frac{\phi (\zeta )\phi (z_1)}{(\zeta z)^2}𝑑\zeta ,$$ so that (4.22) $$g(z)=h_1(z)+h_2(z),$$ we now show that $`h_1_{\mathrm{}}`$ is sufficiently small. Indeed, we obtain by Lemma 4.4 and 4.2 that $`h_1_{\mathrm{}}`$ $``$ $`\underset{z\overline{G}}{\mathrm{max}}{\displaystyle _\gamma ^{}}{\displaystyle \frac{|\phi (\zeta )\phi (z_1)|}{|\zeta z|^2}}|d\zeta |{\displaystyle _\gamma ^{}}{\displaystyle \frac{\mathrm{exp}\left(c/|\zeta z_1|^{p1}\right)}{[d(\zeta ,L)]^2}}|d\zeta |`$ $``$ $`{\displaystyle _\gamma ^{}}{\displaystyle \frac{\mathrm{exp}\left(c[d(\zeta ,L)]^{\frac{p1}{K^2(q1)+1}}\right)}{[d(\zeta ,L)]^2}}|d\zeta |`$ $``$ $`\underset{\zeta \gamma ^{}}{\mathrm{max}}{\displaystyle \frac{\mathrm{exp}\left(c[d(\zeta ,L)]^{\frac{p1}{K^2(q1)+1}}\right)}{[d(\zeta ,L)]^2}}.`$ Since the function $`x^2\mathrm{exp}(cx^a)`$, where $`a,c>0`$, is strictly increasing on an interval $`(0,x_0)`$, we deduce from the previous inequality that (4.23) $$h_1_{\mathrm{}}\frac{\mathrm{exp}\left(c[d(\zeta _R,L)]^{\frac{p1}{K^2(q1)+1}}\right)}{[d(\zeta _R,L)]^2},$$ where $`R=1+2n^s`$ is sufficiently close to 1 and $`\zeta _RL_R`$. It is known that $`\mathrm{\Psi }:=\mathrm{\Phi }^1`$ is Hölder continuous on $`\overline{D^{}}`$ (see Theorem 3 in ), so that $$d(\zeta _R,L):=\underset{tL}{\mathrm{min}}|\zeta _Rt|(R1)^\beta n^{s\beta },$$ for some $`\beta >0.`$ Hence we obtain from (4.23) that (4.24) $$h_1_{\mathrm{}}n^{2s\beta }\mathrm{exp}\left(cn^{\frac{s\beta (p1)}{K^2(q1)+1}}\right),n.$$ Combining (4.21), (4.22) and (4.24), we have $$gp_n_{\mathrm{}}\mathrm{exp}\left(cn^r\right),n,$$ where $`r(0,1)`$ is any number satisfying (4.25) $$r<\mathrm{min}(1s,\frac{s\beta (p1)}{K^2(q1)+1}).$$ Furthermore, this immediately implies that there exists a sequence of polynomials $`\{P_n(z)\}_{n=1}^{\mathrm{}}`$ such that (4.26) $$\phi ^{}P_n_2\phi ^{}P_n_{\mathrm{}}\mathrm{exp}\left(cn^r\right),n,$$ by (4). This concludes the first part of the proof, because we obtain from (4.26) and (1.2) that $`\phi ^{}B_n^{}_2`$ $``$ $`\phi ^{}(P_{n1}P_{n1}(z_0)+1)_2`$ $``$ $`\phi ^{}P_{n1}_2+|1P_{n1}(z_0)|`$ $`=`$ $`\phi ^{}P_{n1}_2+|\phi ^{}(z_0)P_{n1}(z_0)|`$ $``$ $`\mathrm{exp}\left(c(n1)^r\right)\mathrm{exp}\left(cn^r\right),`$ where $`n2.`$ The second part follows from a standard argument on translating the estimate (4.27) $$\phi ^{}B_n^{}_2\mathrm{exp}\left(cn^r\right),n,$$ into (2.1), using the following polynomial inequality (4.28) $$Q_n_{\mathrm{}}n^{P1}Q_n^{}_2,n,$$ which is valid for any $`Q_n(z),`$ with $`Q_n(z_0)=0`$ (see Corollary 2 in ). We write $$\phi (z)=B_n(z)+\underset{k=1}{\overset{\mathrm{}}{}}\left(B_{(k+1)n}(z)B_{kn}(z)\right),zG,$$ so that (4.29) $$\phi B_n_{\mathrm{}}\underset{k=1}{\overset{\mathrm{}}{}}B_{(k+1)n}B_{kn}_{\mathrm{}}.$$ We next estimate terms in the above sum, using (4.28) and (4.27): $`B_{(k+1)n}B_{kn}_{\mathrm{}}`$ $``$ $`\left((k+1)n\right)^{P1}B_{(k+1)n}^{}B_{kn}^{}_2`$ $``$ $`\left((k+1)n\right)^{P1}\left(\phi ^{}B_{kn}^{}_2+B_{(k+1)n}^{}\phi ^{}_2\right)`$ $``$ $`(k+1)^{P1}n^{P1}\mathrm{exp}\left(c(kn)^r\right).`$ It follows from (4.29) that $`\phi B_n_{\mathrm{}}`$ $``$ $`{\displaystyle \underset{k=1}{\overset{\mathrm{}}{}}}(k+1)^{P1}n^{P1}\mathrm{exp}\left(c(kn)^r\right)`$ $`=`$ $`n^{P1}\mathrm{exp}\left(cn^r\right){\displaystyle \underset{k=1}{\overset{\mathrm{}}{}}}(k+1)^{P1}\mathrm{exp}\left(cn^r(k^r1)\right)`$ $``$ $`n^{P1}\mathrm{exp}\left(cn^r\right){\displaystyle \underset{k=1}{\overset{\mathrm{}}{}}}(k+1)^{P1}\mathrm{exp}\left(c(k^r1)\right)`$ $``$ $`n^{P1}\mathrm{exp}\left(cn^r\right).`$ Since we can drop the term $`n^{P1}`$ in the last estimate, by slightly decreasing $`r`$, equation (2.1) is proved. ∎ Proof of Theorem 2.2. We essentially follow Keldysh’s construction in , augmented with Lemma 4.5. Let $`G_1\{|z|<1\}`$ be a symmetric in the real axis domain, which is bounded by a piecewise analytic Jordan curve with the only corner point $`\xi _1(0,1)`$. Clearly, if the inner angle at $`\xi _1`$ is $`\alpha _1\pi `$, where $`\alpha _1(0,1)`$ is irrational, then the conformal mapping of $`G_1`$ onto a disk cannot be analytic in a neighborhood of $`\xi _1`$. Therefore, $`G_1`$ satisfies the assumption of Lemma 4.5 and we can find a point $`\xi _2(\xi _1,1)`$ and a number $`n_1`$ such that $$|B_{n_1,1}(\xi _2)|>2,$$ where $`B_{n,1}(z)`$ is the $`n`$-th Bieberbach polynomial associated with $`G_1`$. Next, we similarly construct a domain $`G_2`$ bounded by a symmetric piecewise analytic curve with the only corner point at $`\xi _2`$, so that $`G_1G_2\{|z|<1\}`$ and $$\underset{|z|1}{\mathrm{max}}|B_{n,1}(z)B_{n,2}(z)|<\frac{1}{2^2},nn_1.$$ This can be always achieved by taking the boundary of $`G_2`$ sufficiently close to the continuum $`\overline{G_1}[\xi _1,\xi _2]`$, because the coefficients of Bieberbach polynomials are rational functions of the moments $`_{G_1}z^k\overline{z}^{\mathrm{}}𝑑x𝑑y`$, by Gram-Schmidt orthonormalization scheme and (1.5), and are continuously dependent on the domain. Proceeding in this fashion, we obtain a sequence of domains $`G_1G_2\mathrm{}G_m\mathrm{}\{|z|<1\}`$, such that (4.30) $$|B_{n_m,m}(\xi _{m+1})>2m,m,$$ and (4.31) $$\underset{|z|1}{\mathrm{max}}|B_{n,m}(z)B_{n,m+1}(z)|<\frac{1}{2^{m+1}},nn_m,$$ where $`B_{n,m}(z)`$ is the $`n`$-th Bieberbach polynomial associated with the domain $`G_m`$. Furthermore, we can carry out this construction in such a way that $`G_m`$ converges to a domain $`G`$, as $`m\mathrm{}`$, which is bounded by a piecewise smooth curve symmetric in the real axis, with the only singular point $`\xi :=lim_m\mathrm{}\xi _m`$. Let $`B_n(z)`$ be the $`n`$-th Bieberbach polynomial for $`G`$. Then we have that (4.32) $$\underset{m\mathrm{}}{lim}B_{n,m}(z)=B_n(z),n,$$ where the convergence is uniform on compact subsets of $``$ for each fixed $`n`$. It follows from (4.31) and (4.32) that $$\underset{|z|1}{\mathrm{max}}|B_{n_m,m}(z)B_{n_m}(z)|<\frac{1}{2^m},m,$$ which implies that $$|B_{n_m}(\xi _{m+1})|>m,m,$$ by (4.30). Hence $$\underset{n\mathrm{}}{lim\; sup}B_n_{\mathrm{}}=\mathrm{}.$$ But this is impossible if $`G`$ has a non-zero angle at $`\xi `$, as Bieberbach polynomials converge uniformly on $`\overline{G}`$, for domains with quasiconformal boundary (see ). Thus, we are forced to conclude that $`G`$ has an outward pointing cusp at $`\xi `$, according to our construction. To show that $`G`$ is analytic outside of any neighborhood of $`\xi `$, we specify the construction of the domains $`G_m`$ as follows. The piecewise analytic boundary of $`G_m`$, with a corner at $`\xi _m`$, is defined by $`G_m=\tau _m([1,1]),\tau _m(1)=\tau _m(1)=\xi _m,`$ for a mapping $`\tau _m`$ analytic in $`\{w:|w|<1+\epsilon _m\},`$ where $`\epsilon _m0,`$ as $`m\mathrm{}.`$ Clearly, each $`\tau _m`$ is bounded in the unit disk $`D`$, for any $`m.`$ Therefore we can find a subsequence $`\tau _{m_k}`$ that converge locally uniformly in $`D`$ to an analytic mapping $`\tau `$, by a normal families argument. It follows from our geometric construction that $`G=\tau ([1,1]),\tau (1)=\tau (1)=\xi .`$ Proof of Remark 2.3. We need to show that an analytic arc, which is different from a segment of the real axis, can only have a finite order of contact with the real axis. It is sufficient to consider the case of an analytic mapping $`\tau :[0,1]\gamma `$, defining the arc $`\gamma `$, such that $$\tau (w)=\underset{n=1}{\overset{\mathrm{}}{}}c_nw^n,$$ where this series converges in a neighborhood of $`w=0.`$ Suppose that $`a_n=\mathrm{}c_n`$ and $`b_n=\mathrm{}c_n`$. Hence $$x(t):=\mathrm{}\tau (t)=\underset{n=1}{\overset{\mathrm{}}{}}a_nt^n\text{ and }y(t):=\mathrm{}\tau (t)=\underset{n=1}{\overset{\mathrm{}}{}}b_nt^n,$$ for $`t[0,\epsilon ).`$ If the arc $`\gamma `$ has higher order of tangency than at any $`x^p`$-type zero angle at $`\tau (0)=0`$, then $$\underset{t0^+}{lim}\frac{y(t)}{(x(t))^p}=0,p.$$ Since $`a_n0`$ for some $`n`$, we obtain that $$\underset{t0^+}{lim}\frac{y(t)}{t^p}=0,p.$$ Consequently, $`b_n=0,n,`$ and $`y(t)=\mathrm{}\tau (t)0`$ for $`t[0,\epsilon ).`$ We are forced to conclude that $`\gamma `$ is a subset of the real axis in a neighborhood of $`w=0`$, which is an obvious contradiction. ∎ Proof of Theorem 3.1. Note that (3.1) follows directly from (4.5) in Lemma 4.4. To prove (3.2), we consider the analytic continuation of the conformal mapping $`\phi `$, constructed in Lemma 4.4. Thus $`\phi `$ is analytic in a larger domain $`\stackrel{~}{G}`$, such that (4.33) $$|\phi (t)\phi (z)|\mathrm{exp}\left(\frac{c}{|tz|^{p1}}\right),t\stackrel{~}{G},|tz|<d,$$ where $`d>0`$ is sufficiently small. Moreover, (4.3) gives the following estimate for the distance from $`t`$ to $`\stackrel{~}{G}`$: (4.34) $$d(t,\stackrel{~}{G})|tz|^a,t\overline{G},0<|tz|<d,$$ where $`a>1.`$ Letting $`t\overline{G},0<|tz|<d`$, we write $$\phi ^{(k)}(t)=\frac{k!}{2\pi i}_{|wt|=r}\frac{\phi (w)dw}{(wt)^{k+1}},k,$$ and estimate by (4.33): $`|\phi ^{(k)}(t)|`$ $`=`$ $`\left|{\displaystyle \frac{k!}{2\pi i}}{\displaystyle _{|wt|=r}}{\displaystyle \frac{\phi (w)dw}{(wt)^{k+1}}}{\displaystyle \frac{k!}{2\pi i}}{\displaystyle _{|wt|=r}}{\displaystyle \frac{\phi (z)dw}{(wt)^{k+1}}}\right|`$ $``$ $`{\displaystyle \frac{k!}{2\pi }}{\displaystyle _{|wt|=r}}{\displaystyle \frac{|\phi (w)\phi (z)||dw|}{r^{k+1}}}{\displaystyle \frac{k!}{r^k}}\mathrm{exp}\left({\displaystyle \frac{c}{|wz|^{p1}}}\right)`$ $``$ $`{\displaystyle \frac{k!}{r^k}}\mathrm{exp}\left({\displaystyle \frac{c}{(r+|tz|)^{p1}}}\right).`$ Observe that we can use any $`r<d(t,\stackrel{~}{G})`$, which implies with $`r|tz|^a`$ that $$|\phi ^{(k)}(t)||tz|^{ka}\mathrm{exp}\left(c/|tz|^{p1}\right),t\overline{G},0<|tz|<d.$$ Hence $$\underset{\genfrac{}{}{0pt}{}{tz}{t\overline{G}}}{lim}\frac{|\phi ^{(k)}(t)|}{|tz|^m}=0,k,m,$$ so that (3.2) follows. ∎ Proof of Corollary 3.2. It is clear that $`\phi `$ is analytic in a neighborhood of every $`zG`$, which is not a cusp point, by the analytic continuation construction of Lemma 4.4. On the other hand, if $`zG`$ is at a cusp, then we let $`\phi ^{(k)}(z)=0,k,`$ so that $`\phi ^{(k)}`$ is continuous at $`z`$ by (3.2) of Theorem 3.1. It follows that $`\phi ^{(k)}C(\overline{G}),k,`$ i.e., (3.3) holds true. ∎
warning/0001/hep-th0001162.html
ar5iv
text
# References Whereas 2-dimensional field-theoretic models like the Schwinger model or the Gross-Neveu model are by now well understood also at finite temperature , the situation is less clear in the case of the ’t Hooft model , large $`N`$ QCD<sub>2</sub> with fundamental quarks. In view of the similarity between these models, notably their chiral aspects, and the vast amount of literature on QCD<sub>2</sub>, this is rather surprising. Briefly, the present situation is as follows: McLerran and Sen argue that there is no deconfining phase transition, except possibly at infinite temperature. Ming Li , using standard finite temperature field theory methods, concludes that chiral symmetry may get restored in the limit $`T\mathrm{}`$. In both of these studies, severe infrared problems were encountered, either in the form of divergent diagrams or ambiguous quark self energies. In a different vein, several studies have addressed QCD<sub>2</sub> on a spatial circle at zero temperature. By interchanging Euclidean time with space and invoking covariance, this can also be reinterpreted as finite temperature calculations for a spatially extended system, even though the techniques used are quite different. Lenz et al. observe a chiral phase transition in the massless ’t Hooft model at some critical length, strongly reminiscent of the Gross-Neveu model. Dhar et al. treat the zero mode gluons in a more ambitious way than ref. , using technology from matrix models and string theory, but are not able to fully solve the resulting complicated equations. They propose that the gauge variables get decompactified by the fermions in complete analogy with the Schwinger model , a claim which has recently been disputed by Engelhardt . In view of this unsatisfactory state of the art, we have reconsidered the ’t Hooft model on a cylinder and, in particular, tried to clarify the role of the zero mode gluons. We shall present here a novel treatment of the leading order in the large $`N`$ expansion, which seems to resolve the above sketched discrepancies. This opens the way to a controlled study of the $`1/N`$ corrections expected to reveal the true, “hadronic” physics of the model. We work canonically on a spatial circle of length $`L`$ in the gauge $`_1A_1=0`$, $`(A_1)_{ij}=\delta _{ij}\frac{\phi _i}{gL}`$ diagonal in color. The Hamiltonian reads (cf. ) $$H=H_\mathrm{g}+H_\mathrm{f}+H_\mathrm{C},$$ (1) with the gauge field kinetic energy $$H_\mathrm{g}=\frac{g^2L}{4}\underset{i}{}\frac{^2}{\phi _i^2},$$ (2) the quark kinetic energy $$H_\mathrm{f}=\underset{n,i}{}\frac{2\pi }{L}\left(n+\frac{\phi _i}{2\pi }\right)\left(a_i^{}(n)a_i(n)b_i^{}(n)b_i(n)\right)+m\underset{n,i}{}\left(a_i^{}(n)b_i(n)+b_i^{}(n)a_i(n)\right),$$ (3) and the Coulomb interaction $$H_\mathrm{C}=\frac{g^2L}{16\pi ^2}\underset{n,i,j}{}\frac{j_{ij}(n)j_{ji}(n)}{\left(n\frac{\phi _j\phi _i}{2\pi }\right)^2}.$$ (4) Here, the $`a_i(n),b_i(n)`$ denote second quantized right- and left-handed quarks, respectively, and the currents $`j_{ij}(n)`$ can be taken in the U($`N`$) form at large $`N`$, $$j_{ij}(n)=\underset{n^{}}{}\left(a_j^{}(n^{})a_i(n^{}+n)+b_j^{}(n^{})b_i(n^{}+n)\right).$$ (5) As explained elsewhere , due to the curved configuration space of the $`\phi _i`$ and the SU($`N`$) Haar measure originally appearing in $`H_\mathrm{g}`$, this Hamiltonian has to be supplemented by the following boundary condition for the wavefunctionals, $$\mathrm{\Psi }(\phi _1,\mathrm{},\phi _N;\mathrm{fermions})=0\mathrm{if}\phi _i=\phi _j\mathrm{mod}2\pi .$$ (6) In ref. , quantization was performed after complete classical gauge fixing. In this way, the fact that the $`\phi _i`$ are curvilinear coordinates is missed. This led to the assumption that all the $`\phi _i`$ are frozen at the value $`\pi `$ in the large $`N`$ limit. In the resulting purely fermionic theory, the only remnant of the gluons are antiperiodic boundary conditions for the quarks in the compact space direction. In the meantime, this whole approach has been put on a more rigorous basis by first quantizing in the Weyl gauge and then resolving the Gauss law quantum mechanically . This made it clear that the $`\phi _i`$ are parameters on the group manifold with corresponding Jacobian, the SU($`N`$) reduced Haar measure. When solving the theory, it is then possible to restrict oneself to the smallest region in field space bounded by zeros of the Jacobian, cf. ref. where the consequences for SU(2), SU(3) where explored in the case of adjoint fermions. How can this be generalized to the large $`N`$ limit? A definite choice of “fundamental domain” obviously means that the $`\phi _i`$ always remain ordered, say $`0\phi _1\phi _2\mathrm{}\phi _N2\pi `$. If we think of the gluons as particles on a circle, they cannot cross each other and become closely packed in the limit $`N\mathrm{}`$. Their fluctuations are suppressed by $`1/N`$, simply due to lack of space. The only degree of freedom left, the collective rotation of this “pearl necklace”, is a U(1) factor which anyway is not present in the SU($`N`$) theory. This suggests that the correct choice for the gluon background field as seen by the fermions is not $`\phi _i=\pi `$, but rather the continuum limit of the lattice points $$\phi _i=2\pi \frac{i}{N},i=1\mathrm{}N.$$ (7) Instead of antiperiodic boundary conditions, the fermions then acquire color dependent boundary conditions which interpolate smoothly between the phases 0 and $`2\pi `$, $$\psi _k(L)=\mathrm{e}^{\mathrm{i2}\pi k/N}\psi _k(0),(k=1,\mathrm{},N).$$ (8) Note that in such a completely gauge fixed formulation, there is nothing wrong with having a color dependence of this type. In the thermodynamic limit $`L\mathrm{}`$, both of these choices of the gluon field configuration, $`\phi _i=\pi `$ or eq. (7), will become indistinguishable and yield the well-known results. We now show that at finite $`L`$, the effects of the gluons on the quarks is radically different in these two cases, and that it is the spread out distribution (7) which is in fact the correct one. The fermions can be treated in a relativistic Hartree-Fock approximation along the lines explained in ref. . In this approach, all the information about the vacuum is encoded in the Bogoliubov angle $`\theta _i(n)`$, related to self-consistent Hartree-Fock spinors via $$u_i(n)=\left(\begin{array}{c}\mathrm{cos}\theta _i(n)/2\\ \mathrm{sin}\theta _i(n)/2\end{array}\right),v_i(n)=\left(\begin{array}{c}\hfill \mathrm{sin}\theta _i(n)/2\\ \hfill \mathrm{cos}\theta _i(n)/2\end{array}\right).$$ (9) The Bogoliubov angles in turn are determined by the gap equation, $$\frac{2\pi }{L}(n+\alpha _i)\mathrm{sin}\theta _i(n)m\mathrm{cos}\theta _i(n)+\frac{g^2L}{16\pi ^2}\underset{n^{},j}{}\frac{\mathrm{sin}\left(\theta _i(n)\theta _j(nn^{})\right)}{(n^{}\alpha _j+\alpha _i)^2}=0.$$ (10) Here, we have switched to the slightly more convenient variable $`\alpha _i=\frac{\phi _i}{2\pi }[0,1]`$ for the gluons. If $`\alpha _i=1/2`$ as chosen in ref. , $`\theta _i(n)`$ becomes $`i`$-independent and we recover the old gap equation considered in that work. Now, we assume $`\alpha _i=i/N`$ and perform the large $`N`$ limit before solving the gap equation. Since $`\alpha _i`$ becomes a continuous variable, we replace $`\theta _i(n)\theta _\alpha (n)`$ and $`_jN_0^1d\alpha ^{}`$, with the result $$\frac{2\pi }{L}(n+\alpha )\mathrm{sin}\theta _\alpha (n)m\mathrm{cos}\theta _\alpha (n)+\frac{Ng^2L}{16\pi ^2}\underset{n^{}}{}_0^1d\alpha ^{}\frac{\mathrm{sin}\left(\theta _\alpha (n)\theta _\alpha ^{}(nn^{})\right)}{(n^{}\alpha ^{}+\alpha )^2}=0.$$ (11) This infinite set of coupled integral equations collapses into a single, one-dimensional integral equation, if we set $$\theta _\alpha (n)=\theta (n+\alpha ).$$ (12) Since $`n`$ is integer and $`\alpha [0,1]`$, this step in effect decompactifies the original spatial circle. With this ansatz, the notation $`n+\alpha =\nu `$, $`nn^{}+\alpha ^{}=\nu ^{}`$ (where $`\nu ,\nu ^{}`$ are dimensionless, continuous variables) and the substitution $`_n^{}_0^1d\alpha ^{}_{\mathrm{}}^{\mathrm{}}d\nu ^{}`$, we obtain $$\frac{2\pi }{L}\nu \mathrm{sin}\theta (\nu )m\mathrm{cos}\theta (\nu )+\frac{Ng^2L}{16\pi ^2}\mathrm{d}\nu ^{}\frac{\mathrm{sin}(\theta (\nu )\theta (\nu ^{}))}{(\nu \nu ^{})^2}=0.$$ (13) Guided by what is known from the ’t Hooft model in the limit $`L\mathrm{}`$, we have defined the integral as principal value integral. After rescaling the variables via $$\frac{2\pi }{L}\nu :=p,\frac{2\pi }{L}\nu ^{}:=p^{},$$ (14) where $`p,p^{}`$ have the dimension of momenta, we recover exactly the continuum version of the Hartree Fock equation, namely $$p\mathrm{sin}\theta \left(\frac{Lp}{2\pi }\right)m\mathrm{cos}\theta \left(\frac{Lp}{2\pi }\right)+\frac{Ng^2}{4}\frac{\mathrm{d}p^{}}{2\pi }\frac{\mathrm{sin}\left(\theta \left(\frac{Lp}{2\pi }\right)\theta \left(\frac{Lp^{}}{2\pi }\right)\right)}{(pp^{})^2}=0.$$ (15) Denoting the Bogoliubov angle of the continuum ’t Hooft model by $`\theta _{\mathrm{cont}}(p)`$, we conclude that $$\theta (\nu )=\theta _{\mathrm{cont}}\left(\frac{2\pi }{L}\nu \right),$$ (16) or, in terms of the original, color dependent Bogoliubov angle, $$\theta _i(n)\theta _{\mathrm{cont}}\left(\frac{2\pi }{L}\left(n+\frac{i}{N}\right)\right),(N\mathrm{}).$$ (17) This last relation becomes exact in the limit $`N\mathrm{}`$ only. In this limit, the color- and $`L`$-dependent Bogoliubov angles for the ’t Hooft model on the circle of length $`L`$ are all given by one universal function, namely the momentum dependent Bogoliubov angle of the ’t Hooft model on the infinite line. We emphasize that this universality only holds for the “pearl necklace” type distribution of gauge variables, eq. (7). If the $`\phi _i`$ are all set equal to $`\pi `$, there is no analytic way known how to relate $`\theta (n)`$ for different $`L`$ values, but one has to solve the gap equation numerically for each $`L`$ . The upshot of this simple exercise is the following: In the large $`N`$ limit, the gluon variables influence the fermion boundary conditions in such a way that the circle gets replaced by a line; they decompactify space-time. The length $`L`$ of the spatial circle becomes an irrelevant parameter. To confirm this last point, let us evaluate those ground state expectation values which are of interest for the bulk thermodynamic properties of the system, if one interchanges $`L`$ and $`\beta =1/T`$. This can be done most conveniently with the help of the key relations $`a_i^{}(n)a_i(n^{})`$ $`=`$ $`{\displaystyle \frac{1}{2}}\delta _{n,n^{}}\left(1\mathrm{cos}\theta _i(n)\right),`$ $`b_i^{}(n)b_i(n^{})`$ $`=`$ $`{\displaystyle \frac{1}{2}}\delta _{n,n^{}}\left(1+\mathrm{cos}\theta _i(n)\right),`$ $`a_i^{}(n)b_i(n^{})`$ $`=`$ $`b_i^{}(n)a_i(n^{})={\displaystyle \frac{1}{2}}\delta _{n,n^{}}\mathrm{sin}\theta _i(n).`$ (18) The vacuum energy density is given by $`_{\mathrm{vac}}`$ $`=`$ $`{\displaystyle \frac{1}{L}}{\displaystyle \underset{n,i}{}}\left({\displaystyle \frac{2\pi }{L}}(n+\alpha _i)\mathrm{cos}\theta _i(n)+m\mathrm{sin}\theta _i(n)\right)`$ (19) $`+{\displaystyle \frac{g^2}{32\pi ^2}}{\displaystyle \underset{n,n^{},i,j}{}}{\displaystyle \frac{1\mathrm{cos}(\theta _i(n)\theta _j(nn^{})}{(n^{}\alpha _j+\alpha _i)^2}}.`$ It corresponds to the negative of the pressure, in the other picture. Using the same substitutions as above, $`_{\mathrm{vac}}`$ can be converted into the standard continuum expression, $`_{\mathrm{vac}}`$ $`=`$ $`{\displaystyle \frac{N}{L}}{\displaystyle \underset{n}{}}{\displaystyle _0^1}d\alpha \left({\displaystyle \frac{2\pi }{L}}(n+\alpha )\mathrm{cos}\theta (n+\alpha )+m\mathrm{sin}\theta (n+\alpha )\right)`$ (20) $`+{\displaystyle \frac{N^2g^2}{32\pi ^2}}{\displaystyle _0^1}d\alpha {\displaystyle _0^1}d\alpha ^{}{\displaystyle \underset{n,n^{}}{}}{\displaystyle \frac{1\mathrm{cos}(\theta (n+\alpha )\theta (nn^{}+\alpha ^{}))}{(n^{}\alpha ^{}+\alpha )^2}}`$ $`=`$ $`N{\displaystyle \frac{\mathrm{d}p}{2\pi }\left(p\mathrm{cos}\theta _{\mathrm{cont}}(p)+m\mathrm{sin}\theta _{\mathrm{cont}}(p)\right)}`$ $`+{\displaystyle \frac{N^2g^2}{8}}{\displaystyle \frac{\mathrm{d}p}{2\pi }\frac{\mathrm{d}p^{}}{2\pi }\frac{1\mathrm{cos}(\theta _{\mathrm{cont}}(p)\theta _{\mathrm{cont}}(p^{}))}{(pp^{})^2}}.`$ Since the right hand side is $`L`$-independent, the pressure will be equal to the vacuum pressure at all temperatures, to leading order in $`1/N`$. Hence, there is no observable pressure of order $`N`$, as one would expect from a confined system of quarks. Treating the quark condensate along the same lines, we find $`\overline{\psi }\psi `$ $`=`$ $`{\displaystyle \frac{1}{L}}{\displaystyle \underset{n,i}{}}\mathrm{sin}\theta _i(n)`$ (21) $`=`$ $`{\displaystyle \frac{N}{L}}{\displaystyle \underset{n}{}}{\displaystyle _0^1}d\alpha \mathrm{sin}\theta (n+\alpha )`$ $`=`$ $`N{\displaystyle \frac{\mathrm{d}p}{2\pi }\mathrm{sin}\theta _{\mathrm{cont}}(p)}.`$ Once again, the sum over the discrete momenta and the color sum in the large $`N`$ limit conspire to produce the continuum result, independently of the starting $`L`$ value. In the alternative thermodynamic view, the condensate does not depend on temperature to leading order in $`1/N`$. This leaves no room for a chiral phase transition, not even in the limit $`T\mathrm{}`$. Finally, consider the expectation value of the Polyakov loop. Due to our classical treatment and the assumed field configuration, we trivially get zero, $$𝒫=\frac{1}{N}\underset{i}{}\mathrm{e}^{\mathrm{i}\phi _i}_0^{2\pi }\frac{\mathrm{d}\phi }{2\pi }\mathrm{e}^{\mathrm{i}\phi }=0.$$ (22) After interchanging space and time, this signals confinement of static charges at any temperature . Physically, one would expect screening by dynamical quarks, but this cannot be seen yet in leading order in the $`1/N`$ expansion. In the Wilson loop calculation for instance, screening by dynamical quarks involves at least one additional fermion loop. As is well known, such diagrams are suppressed as compared to planar gluonic diagrams by a factor $`1/N`$. A similar argument holds for the Polyakov loop. Notice that so far, we have discussed the influence of the gauge fields on the quarks which is indeed dramatic. Vice versa, we do not expect the quarks to influence significantly the gluon zero point motion, again due to the effects of the Jacobian. The kinetic energy $`H_\mathrm{g}`$ will give the same result as in pure Yang Mills theory. Since this contribution to the energy density is $`L`$-independent, it again yields zero pressure, reflecting the absence of physical gluonic excitations in 1+1 dimensions. We now comment on the various studies of the finite temperature ’t Hooft model mentioned in the beginning. The mechanism which we propose here has some similarity with the color singlet projection in the partition function as discussed by McLerran and Sen , in particular concerning the $`N`$-dependence of the thermodynamic potential. In Ming Li’s calculation , the problem seems to be the quark self energy. In Ref. , it was found that the principal value regulated self energy should be supplemented by an additive constant $`Ng^2L/48`$ which diverges in the limit $`L\mathrm{}`$. This constant drops out of the calculation of color singlet mesons, but cannot be simply ignored at the Hartree Fock level (the principal value prescription makes sense for a first order pole, but not for a second order pole). If we included this infinite constant into Ming Li’s calculation (or, equivalently, a finite temperature Hartree Fock calculation), all the thermal factors would trivially vanish and we would also get zero pressure and a $`T`$-independent condensate. In this sense, the finite $`L`$ calculation presented above and the conventional thermodynamic calculation are fully consistent. Similar arguments have already been put forward heuristically in ref. . As discussed above, the problem with the chiral phase transition seen in ref. is the neglect of Jacobian and consequently freezing of gluons at the point $`\phi _i=\pi `$. The finite $`L`$ version of the theory solved there is apparently not the gauge theory QCD<sub>2</sub>, but has to be interpreted as some other interacting fermion theory of Gross-Neveu type with a Coulomb potential instead of a contact interaction. It cannot do full justice to the confinement of quarks. Finally, in , it has tacitly been assumed that the Jacobian can be accounted for by treating gluons as non-relativistic fermions. This is certainly true for pure Yang Mills theory where the idea originally came up . In the presence of quarks, the Hamiltonian is not invariant under permutations of the $`\phi _i`$. As can be seen from eqs. (14), it is only invariant if one simultaneously permutes the quark colors — this is a residual gauge transformation. Antisymmetrization of the gluon variables alone is thus not compatible with the time evolution of the system and cannot be used to satisfy the boundary condition (6). The other point in which we differ from ref. is the issue of decompactification: In the Schwinger model on the circle, the pure gauge field is a particle on a circle. This circle gets decompactified by the fermions — the axial charge provides the missing integer part of the coordinate . For reasons discussed in , such a phenomenon is hard to conceive in the SU($`N`$) case, unlike what has been put forward in . Interestingly, we have found exactly the opposite effect: The gluons decompactify the spatial circle on which the fermions live, just because they are so much constrained by their own, compact configuration space. This seems to be the mechanism by which a confined system of quarks in 1+1 dimensions avoids wrong thermodynamic behaviour to leading order in $`1/N`$. We should like to thank M. Engelhardt for a helpful discussion and M. Shifman for his interest in our work.
warning/0001/hep-lat0001017.html
ar5iv
text
# General Algorithm For Improved Lattice Actions on Parallel Computing Architectures ## I Introduction It is almost universally accepted that Quantum Chromodynamics (QCD) is the underlying quantum field theory of the strong interaction , which binds atomic nuclei and fuels the sun and the stars. Strongly interacting particles are referred to as hadrons, which include for example protons and neutrons that make up atomic nuclei as well as a wide variety of particles that are produced in particle accelerators and from astrophysical sources. These hadrons are made up of quarks and gluons, which are the underlying constituents in QCD. The quarks are spin-1/2 particles (i.e., fermions) and the gluons are massless spin-1 particles (i.e., gauge bosons). The quarks interact strongly through their “colour” charge through the exchange of gluons. The 8 gluons of SU(3) (i.e., one for each generator of SU(3)) themselves carry colour and hence interact with themselves as well as with the quarks. This is the essential difference between QCD and the corresponding theory of photons and electrons referred to as quantum electrodynamics (QED) and has far reaching consequences since the theories have entirely different behavior. The are very few first-principles methods for studying QCD in the nonperturbative low-energy regime. The most widely used of these is the so-called Lagrangian-based lattice field theory, which formulates the field theory on a space-time lattice . An alternative lattice approach is based on the Hamiltonian formulation of quantum field theory and makes use of cluster decompositions and again Monte Carlo methods to carry out the simulations. In addition, there are numerous studies based on a light-front formulation of QCD and much use has been made of Schwinger-Dyson equations to assist with the construction of QCD-based quark models. The Lagrangian-based lattice technique simulates the functional integral using a four-dimensional hypercubic Euclidean spacetime lattice together with Monte Carlo methods for generating an ensemble of gluon field configurations with the appropriate Boltzmann distribution $`\mathrm{exp}(S_G)`$, where $`S_G`$ is a discretized form of the QCD gluon action on the hypercubic lattice. The simplest discretizations of the QCD action involve only nearest neighbours on the lattice and have $`𝒪(a^2)`$ errors, where $`a`$ is the lattice spacing. Improved actions represent a major advance for the field of lattice gauge theory, where by using increasingly non-local discretizations of the QCD action we can obtain the same accuracy with far fewer lattice points and hence far less computational time and effort. The purpose of the present work is to describe an algorithm which allows us to implement an arbitrarily improved (i.e., arbitrarily non-local) action in an efficient way. For further details on the state of the art lattice QCD techniques see for example Ref. . Another related and equally important advance is the technique of nonperturbative improvement (e.g., mean-field improvement) which corrects for some of the major nonperturbative effects (the so-called tadpole contributions) and hence more quickly brings the lattice results to their continuum form by improving the matching with perturbation theory at a given lattice spacing $`a`$ . It is the combination of improved actions and nonperturbative improvement that together have come to represent a significant advance for the field . Lattice QCD is based on a Monte Carlo treatment of the path integral formulation, which makes it a computationally demanding method for calculating physical observables. The gluon field is represented by $`3\times 3`$ complex $`SU(3)`$ matrices, where there is one such $`SU(3)`$ matrix associated with every link on the lattice. The links lie only along one of the four Cartesian directions and join neighbouring lattice sites. Since all lattice links require identical numerical calculations, lattice gauge theory is ideally suited for parallel computers. There are various types of improved actions and, as explained above, these are all based on the idea of eliminating the discretization errors that occur when passing from continuum physics to the discretized lattice version. The simplest (i.e., non-improved) gluon action is the so called standard Wilson action and consists of $`1\times 1`$ Wilson loops or as they are frequently called plaquettes. We shall often refer to the Wilson loops used to build up lattice actions as plaquettes. The need to build the gluon action out of closed loops arises from the need to maintain exact $`SU(3)`$ gauge invariance in the discrete lattice action. This $`1\times 1`$ loop action was first proposed by Wilson in the early 70’s and has been used extensively over the years. It consists of taking an arbitrary starting site, say $`x`$, on the lattice and stepping around a $`1\times 1`$ loop until returning to the starting point $`x`$. The $`1\times 1`$ Wilson loop is illustrated in Fig. 1. Improving the standard Wilson action is achieved by making use of larger loops (e.g., $`1\times 2`$, $`2\times 2`$, etc.) in the lattice gluon action to eliminate finite lattice spacing artifacts to a given order in $`a^2`$ . For an elegant and detailed discussion of these topics see Ref. . In this article we present an efficient and completely general algorithm that permits one to calculate any improved planar lattice action at any desired level of improvement. By “planar” here we mean that we will consider actions containing two-dimensional loops of arbitrary size which lie in any of the Cartesian planes, (i.e., the $`xy`$, $`xz`$, $`xt`$, $`yz`$, $`yt`$, or $`zt`$ plane). This algorithm has been used in a wide variety of improved action lattice simulations to date . For example, it has been used in studies of the topological structure of the QCD vacuum and the calibration of the various cooling and smearing techniques , the study of discretization errors in the Landau gauge on the lattice , and studies of the static quark potential . It is currently being used in studies of the gluon propagator and highly improved actions . In Sec. II, we briefly describe the tree–level improved action that we have been using in our calculations with our algorithm on a Thinking Machines CM-5. Sec. III gives two possible ways of using the technique for the standard Wilson action. The form of the algorithm appropriate for the first level of improvement (i.e., involving a combination of the elementary square $`1\times 1`$ plaquette and the rectangular $`1\times 2`$ plaquette) is given in Sec. IV. Then in Sec. V we present the general algorithm suitable for an arbitrarily non-local action, (i.e., for an $`n\times m`$ Wilson loop with $`n`$ and $`m`$ being arbitrary positive integers). Sec. VI addresses non-planar issues encountered in nesting specific planar actions. Actions involving non-planar loops are also addressed. Finally, in Sec. VII we present our summary and conclusions. ## II Gauge Action, Masking, and Parallel Computing ### A Lattice Gauge Action for Colour $`SU(3)`$ The standard Wilson action for the gluons is given by $$S_G^\mathrm{W}=\beta \underset{\mathrm{sq}}{}\left[1\frac{1}{3}e\mathrm{Tr}U_{\mathrm{sq}}(x)\right]$$ (1) and a simple tree–level $`𝒪(a^2)`$–improved action (i.e., the action with the first level of improvement) is defined as $$S_G=\frac{5\beta }{3}\underset{\mathrm{sq}}{}e\mathrm{Tr}(1U_{\mathrm{sq}}(x))\frac{\beta }{12u_0^2}\underset{\mathrm{rect}}{}e\mathrm{Tr}(1U_{\mathrm{rect}}(x)).$$ (2) The $`1\times 1`$ square (or plaquette) $`U_{\mathrm{sq}}(x)`$ and the $`1\times 2`$ rectangle $`U_{\mathrm{rect}}(x)`$ are defined by $`U_{\mathrm{sq}}(x)`$ $`=`$ $`U_\mu (x)U_\nu (x+\widehat{\mu })U_\mu ^{}(x+\widehat{\nu })U_\nu ^{}(x)`$ (3) $`U_{\mathrm{rect}}(x)`$ $`=`$ $`U_\mu (x)U_\nu (x+\widehat{\mu })U_\nu (x+\widehat{\nu }+\widehat{\mu })U_\mu ^{}(x+2\widehat{\nu })U_\nu ^{}(x+\widehat{\nu })U_\nu ^{}(x)`$ (4) $`+`$ $`U_\mu (x)U_\mu (x+\widehat{\mu })U_\nu (x+2\widehat{\mu })U_\mu ^{}(x+\widehat{\mu }+\widehat{\nu })U_\mu ^{}(x+\widehat{\nu })U_\nu ^{}(x).`$ (5) Here the variables $`\mu `$ and $`\nu `$ are the direction in which the links are pointing inside the lattice space. There are four directions for a four-dimensional hypercubic lattice. The link product $`U_{\mathrm{rect}}(x)`$ denotes the rectangular $`1\times 2`$ plaquettes and $`u_0`$ is the tadpole improvement factor, commonly known as the mean–field improvement factor which largely corrects for quantum renormalization of the links. In our numerical studies we have typically employed the plaquette definition of the mean–field improvement factor $$u_0=\left(\frac{1}{3}e\mathrm{Tr}U_{\mathrm{sq}}\right)^{\frac{1}{4}}.$$ (6) For the improved action in Eq. (2) the residual perturbative corrections after mean–field improvement are estimated to be of the order of two to three percent . Of course, both Eqs. (1) and (2) reproduce the continuum gluon action as $`a0`$, where $`\beta 6/g^2`$ and $`g`$ is the QCD coupling constant at the scale $`a`$. It is useful to note that our $`\beta =6/g^2`$ differs from the convention of Refs. . A multiplication of our $`\beta `$ in Eq. (2) by a factor of $`5/3`$ reproduces their definition. Let us comment on the lattice configurations that we have generated with the general algorithm described here and which we have used extensively in Refs. . The gauge configurations are generated using the Cabbibo-Marinari pseudo–heat–bath algorithm with three diagonal $`SU_c(2)`$ subgroups. All calculations are performed using a highly parallel code written in CM-Fortran and run on a Thinking Machines Corporations (TMC) CM-5 with appropriate link partitioning. For the standard Wilson action we partition the link variable in a checkerboard fashion. While all calculations to date have been for $`SU(3)`$, there is no restriction in the algorithm on the number of colours for the gauge group and we could just has easily have treated the case of $`SU(N)`$. The mean–field improvement factor was updated on a regular basis during the simulation. Once the lattice is thermalized from a cold start, (after at least five thousand sweeps), the $`u_0`$ factor is held fixed during the generation of the ensemble of gauge field configurations. The ensemble is built up by sampling the fields with a separation of at least 500 Monte Carlo sweeps over the entire lattice to ensure that they are sufficiently decorrelated. For the case of the standard Wilson action, configurations have been generated on a $`16^3\times 32`$ lattice at $`\beta =5.70`$ and a $`24^3\times 36`$ lattice at $`\beta =6.00`$. For the improved action of Eq. (2) we have generated $`8^3\times 16`$, $`12^3\times 24`$, $`16^3\times 32`$, and $`24^3\times 36`$ lattices with $`\beta `$ values of 3.57, 4.10, 4.38, and 5.00 respectively. ### B Masking and Parallel Computing When performing a Monte Carlo sweep of the entire lattice each lattice link must be updated individually using the particular gluon action of interest (e.g., $`S_G`$). The action is used in the Monte Carlo accept/reject step for that link in order that detailed balance is ensured at each link update and hence that it is ensured throughout the entire lattice sweep. It is the combination of randomness in the link updates, the maintenance of detailed balance, and decorrelation (ensured by large sweep numbers between the taking of samples) that ensures the desired ensemble of gauge field configurations are produced with the Boltzmann distribution $`\mathrm{exp}(S_G)`$. In the most naive procedure we move through each link on the lattice consecutively updating them one at a time until we have completed a “sweep” through the entire lattice. We then repeat these lattice sweeps as often as required. This simple procedure is highly inefficient on a parallel computing architecture, where we can be updating many links at the same time. However, there is a fundamental limitation to this parallelism, i.e., we will violate detailed balance and corrupt our data if we try to simultaneously update a link while information about that link is being used in the update of another link. It is crucial that we identify which links can be updated simultaneously and this is determined by the degree of non locality in the action. For example, for an action which contains only nearest neighbor interactions of the links, such as the Wilson action, we can use an efficient “checkerboard” algorithm, which will be described below. In general, the more non local is the lattice gluon action the fewer are the links that can be simultaneously updated. We see that the improvement program is therefore more expensive to implement, but the benefit of improved actions far outweighs this drawback. In order to facilitate our discussions we will refer to the concept of “masking”, where the lattice links not eliminated by the mask are the ones that can be simultaneously updated in a parallel computing environment. The number of independent masks needed for a particular action determines an upper limit to the parallelism that can be used in a single lattice sweep. As we will see, the best that can be done is to have two masks per link direction and this is for the case of nearest neighbor interactions only. We will simplify the presentation in the usual way by rescaling all dimensionful quantities by the lattice spacing $`a`$. This is equivalent to setting $`a=1`$. ## III Masking for the Standard Wilson Action. In the standard Wilson action, where only neighbouring links are connected by the action, we need only two masks for each of the four link directions. There are two different ways of implementing this masking as we will now discuss. ### A Checker Board Masking. The standard Wilson action only involves $`1\times 1`$ Wilson loops (depicted in Fig. 1) and is the most fundamental lattice gluonic action. Whenever a given link is being updated, we must not be attempting to update any of the links within any of the $`1\times 1`$ plaquettes which contains the given link. Consider the link from the lattice site $`x`$ to $`x+\mu `$, where $`\mu `$ is one of the four Cartesian unit vectors $`\widehat{x},\widehat{y},\widehat{z}`$, or $`\widehat{t}`$. We see then that the plaquette in Fig. 1 forms a “staple” consisting of three links in the $`\mu `$-$`\nu `$ plane which is attached to the link of interest $`U_\mu (x)`$. \[Note that we are sometimes using $`x`$ as a shorthand notation for the space-time lattice point $`x^\mu (x,y,z,t)`$ as well as for the $`x`$-coordinate on the $`\widehat{x}`$ axis. The meaning should be clear from the context.\] We could equally well consider the plaquette and staple below the link $`U_\mu (x)`$ in the figure, which also lies in the $`\mu `$-$`\nu `$ plane. In addition, for a given Cartesian direction $`\mu `$, there are three possible choices for $`\nu `$, i.e., there are three orthogonal planes which contain the link and two staples per plane. Let us consider, for example, all of the links in the $`\widehat{x}`$-$`\widehat{y}`$ plane which are oriented in the $`\widehat{x}`$ direction. We can see from Fig. 2 that we can choose a “checkerboard” of such links that can be updated at the same time without interfering with each other. These links are indicated in the figure as highlighted links with arrows. It is easy to see that none of the links to be updated lie in any of the staples for the other links to be updated and that exactly half of the $`\widehat{x}`$-oriented links in this plane can be simultaneously updated at one time. We have identified one of the lattice sites in Fig. 2 as the site $`x`$. If the the link variable $`U_{\widehat{x}}(x)`$ is to updated then from Fig. 2, it is observed that the link variables in the $`\widehat{x}`$ direction that can be simultaneously updated are $`U_{\widehat{x}}(x+2\widehat{x})`$, $`U_{\widehat{x}}(x+4\widehat{x})`$ and so on. So every second link along the $`\widehat{x}`$ direction can be updated at the same time. Now let us consider stepping in the $`\widehat{y}`$ direction. We again see that every second link in that direction can be simultaneously updated. By symmetry the same must also be true for the $`\widehat{z}`$ and $`\widehat{t}`$ directions as depicted in Fig. 3, where we have used a broken dash-dot line to try to indicate the fourth dimension (i.e., for the links that lie in the $`\widehat{x}`$$`\widehat{t}`$ plane). We see that for the link pointing in the $`\widehat{x}`$ direction, the plaquettes (and staples) in the $`\widehat{x}`$$`\widehat{y}`$, $`\widehat{x}`$$`\widehat{z}`$, $`\widehat{x}`$$`\widehat{t}`$ planes are all related by simple rotations about the link. Thus we see that we have now built up a four-dimensional mask for determining which links pointing in the $`\widehat{x}`$ direction can be simultaneously updated. Let us introduce some convenient shorthand notation. If for a given link pointing in the direction $`\mu `$, we must take $`n`$ steps in the direction $`\nu `$ before reaching the next updatable link pointing in the direction $`\mu `$, we will use the notation $`\mu :\nu n\nu `$. For our checkerboard masking we see that for a link pointing in the direction $`\widehat{x}`$ we have to take two steps in each of the Cartesian directions before reaching the next updatable link. Hence we write $`\widehat{x}:\widehat{x}2\widehat{x},\widehat{y}2\widehat{y},\widehat{z}2\widehat{z}`$ $`\mathrm{and}`$ $`\widehat{t}2\widehat{t}.`$ (7) We immediately see that this is also true for links oriented in the $`\widehat{y}`$, $`\widehat{z}`$, and $`\widehat{t}`$ directions so that $`\widehat{y}:\widehat{x}2\widehat{x},\widehat{y}2\widehat{y},\widehat{z}2\widehat{z}`$ $`\mathrm{and}`$ $`\widehat{t}2\widehat{t},`$ (8) $`\widehat{z}:\widehat{x}2\widehat{x},\widehat{y}2\widehat{y},\widehat{z}2\widehat{z}`$ $`\mathrm{and}`$ $`\widehat{t}2\widehat{t},`$ (9) $`\widehat{t}:\widehat{x}2\widehat{x},\widehat{y}2\widehat{y},\widehat{z}2\widehat{z}`$ $`\mathrm{and}`$ $`\widehat{t}2\widehat{t}.`$ (10) Finally, note that when we wish to update all of the links pointing in any one of the four Cartesian directions, say $`\mu `$, we need only two four-dimensional masks. This is because exactly half of the $`\mu `$-oriented links across the entire lattice are considered in each four-dimensional mask. To appreciate this we simply note that for any one of the Cartesian directions one mask can be turned into the checkerboard complement mask for that direction by shifting the mask by one step in any Cartesian direction, (see Fig. 2). So to update all of the links on the lattice we need a total of 8 four-dimensional masks, i.e., 2 masks for each of the four Cartesian directions. In other words, no matter how many nodes we have available on our parallel computing architecture a full lattice updating sweep will require 8 serial masked sweeps to complete with a nearest neighbour action (such as the Wilson action) and checkerboard masking. This is the conventional procedure for the standard Wilson action in lattice QCD studies. In closing this section on the standard Wilson action let us observe in Sec. III B that there is an alternative and equally good “linear” masking for this case. ### B Linear Masking. As an alternative approach to the checker board masking described in Sec. III A, one could partition the links over the lattice in a linear fashion as shown in Fig. 4. If the link variable of interest is $`U_{\widehat{x}}(x)`$ then the next possible link variable in the $`\widehat{x}`$ direction which can be updated is the $`U_{\widehat{x}}(x+\widehat{x})`$ link and then the $`U_{\widehat{x}}(x+2\widehat{x})`$ and so on. We see that all the links on the $`\widehat{x}`$ line can be updated at the same time, since none of these links are contained in the $`1\times 1`$ plaquettes for the other links in the line. Hence we have $`\widehat{x}:\widehat{x}1\widehat{x}`$. Now looking in the $`\widehat{y}`$ direction, we realize that we cannot touch the $`U_{\widehat{x}}(x+\widehat{y})`$ link because it is part of the Wilson loop containing the link variable $`U_{\widehat{x}}(x)`$ which is being updated simultaneously. However, the links $`U_{\widehat{x}}(x+2\widehat{y})`$, $`U_{\widehat{x}}(x+4\widehat{y})`$, etc. can be updated. Consequently, we have $`\widehat{x}:\widehat{y}2\widehat{y}`$ and similarly for steps in the $`\widehat{z}`$ and $`\widehat{t}`$ directions. For a link variable pointing in the $`\widehat{x}`$ direction we then have that $`\widehat{x}:\widehat{x}1\widehat{x},\widehat{y}2\widehat{y},\widehat{z}2\widehat{z}`$ $`\mathrm{and}`$ $`\widehat{t}2\widehat{t}.`$ (11) When the links to be updated are pointing in the other three directions we have $`\widehat{y}:\widehat{x}2\widehat{x},\widehat{y}1\widehat{y},\widehat{z}2\widehat{z}`$ $`\mathrm{and}`$ $`\widehat{t}2\widehat{t},`$ (12) $`\widehat{z}:\widehat{x}2\widehat{x},\widehat{y}2\widehat{y},\widehat{z}1\widehat{z}`$ $`\mathrm{and}`$ $`\widehat{t}2\widehat{t},`$ (13) $`\widehat{t}:\widehat{x}2\widehat{x},\widehat{y}2\widehat{y},\widehat{z}2\widehat{z}`$ $`\mathrm{and}`$ $`\widehat{t}1\widehat{t},`$ (14) for the $`\widehat{y},\widehat{z}`$ and $`\widehat{t}`$ directions respectively. Again, we see that there are two complementary linear masks for links pointing in any given Cartesian direction $`\mu `$. One mask can be obtained from the other by a shift of one step in any of the three Cartesian directions orthogonal to $`\mu `$ as can be appreciated from Fig. 4. Thus this linear masking is equally as efficient as the checkerboard masking of the previous section, since there are 2 masks for each of the 4 Cartesian directions giving a total of 8 masks. ## IV Masking an Improved Action. In this section, we describe the necessary masking procedure for a first-level improved action involving $`1\times 1`$ and $`1\times 2`$ Wilson loops. In particular, in this section we are describing the masking suitable for the improved gauge action of Eq. (2), which has been used extensively by us . Let us again begin by considering the link variable beginning at some lattice site $`x`$ and pointing in the $`\widehat{x}`$ direction, i.e., $`U_{\widehat{x}}(x)`$. We now need to consider both Fig. 3 for the elementary $`1\times 1`$ square plaquette and Fig. 5 for the $`1\times 2`$ rectangular plaquette. In Fig. 5 we show all of the $`1\times 2`$ rectangular plaquettes which contain the link $`U_{\widehat{x}}(x)`$, which is shown as the highlighted horizontal link in the three parts of this figure. Visualizing a four dimensional object on a flat piece of paper can be, to a certain extent, an artistic challenge and so we have again used a dash–dot line to indicate links lying in the $`\widehat{x}`$$`\widehat{t}`$ plane. There are three distinguishable ways to include this link in a $`1\times 2`$ plaquette (the three parts of the figure) and for each of these there are two (mirror-image) rectangles per Cartesian plane and four Cartesian planes. All links in Figs. 3 and 5 with arrows (other than the link $`U_{\widehat{x}}(x)`$ itself) must be omitted from the mask when updating this link with our improved action. We see that there are many excluded links. In Fig. 6 we show which links can be simultaneously updated with the link $`U_{\widehat{x}}(x)`$. We can immediately write down by inspection from this figure that $`\widehat{x}:\widehat{x}2\widehat{x},\widehat{y}3\widehat{y},\widehat{z}3\widehat{z}`$ $`\mathrm{and}`$ $`\widehat{t}3\widehat{t}.`$ (15) This follows since the $`\widehat{z}`$ and $`\widehat{t}`$ cases are identical to the $`\widehat{y}`$ case for this $`\widehat{x}`$–oriented link. The generalization to the other orientations of the links to be updated is straightforward by symmetry $`\widehat{y}:\widehat{x}3\widehat{x},\widehat{y}2\widehat{y},\widehat{z}3\widehat{z}`$ $`\mathrm{and}`$ $`\widehat{t}3\widehat{t},`$ (16) $`\widehat{z}:\widehat{x}3\widehat{x},\widehat{y}3\widehat{y},\widehat{z}2\widehat{z}`$ $`\mathrm{and}`$ $`\widehat{t}3\widehat{t},`$ (17) $`\widehat{t}:\widehat{x}3\widehat{x},\widehat{y}3\widehat{y},\widehat{z}3\widehat{z}`$ $`\mathrm{and}`$ $`\widehat{t}2\widehat{t}.`$ (18) Let us return to the particular case of the masking for $`\widehat{x}`$-oriented links. From Eq. (15) we see that there is symmetry between the $`\widehat{y}`$, $`\widehat{z}`$, and $`\widehat{t}`$ directions and so we will begin by constructing suitable masks for any given equal-$`x`$ hyper-plane, i.e., for the three dimensional space spanned by the unit vectors $`\widehat{y}`$, $`\widehat{z}`$, and $`\widehat{t}`$. Before attempting this, let us first consider Fig. 6 and extend this to three dimensions by imagining that the $`\widehat{z}`$-axis is pointing directly out from the page. We shall temporarily neglect the $`\widehat{t}`$ direction, which is equivalent to simply taking a slice of the four-dimensional lattice with the same value of $`t`$, (i.e., an equal-$`t`$ hyper-plane). Now let us view this three-dimensional lattice by looking along the $`\widehat{x}`$-axis at one particular equal-$`x`$ plane. We will then be presented with end-views of updatable links in the $`\widehat{y}`$$`\widehat{z}`$ plane. For every fixed value of $`z`$ there are three different masks needed for $`y`$ and vice versa. Also, there is no restriction on simultaneously updating diagonally shifted links, since we are only considering planar actions at this point. It is not difficult to see that we can cover all of the nine lattice links that need to be updated with three orthogonal masks as shown in Fig. 7. In this figure $`x`$-oriented links which can be updated at the same time are indicated by a solid dot. Note that each of these masks is related by a diagonal shift of the nine-point lattice “window”. We can now also extend this thinking to include the $`t`$ direction, by stacking the three two-dimensional $`y`$$`z`$ masks on top of each other as shown in Fig. 8. We must stack the planes so that when viewed along any of the three axes the solid dots in any one Cartesian planes always have the appearance of one of the planes in Fig. 7. We see that this can be achieved in three ways by the stacking in Fig. 8 and its two cyclic permutations. These three three-dimensional masks when summed give the identity (i.e., the sum includes all points) and are orthogonal to each other (i.e., the sum includes all points only once). We can now give a simple geometrical picture of what we are doing, which will simplify the generalization that we give in the next section. For $`\widehat{x}`$-oriented links, the directions $`\widehat{y}`$, $`\widehat{z}`$, and $`\widehat{t}`$ directions are all symmetrical and each direction requires a step of 3 to reach the next updatable link. Hence, we need to construct a complete set of orthogonal masks in three dimensions for a $`3\times 3\times 3`$ cube, where no two points in the cube lie on the same Cartesian axis (i.e., only diagonally related points). This is simple to do. Let us consider the bottom plane (i.e., plane 1) of Fig. 8 and connect the three solid dots by a diagonal line. We see that plane 2 is obtained from plane 1 by a diagonal shift of this line by one diagonal half-step, and similarly for plane 3. In visualizing this it may help to imagine surrounding the cube by many identical copies of itself and moving the diagonal line through diagonal half-steps across all of these cubes simultaneously. All three three-dimensional masks are obtained in the same way but start with plane 1, plane 2, and plane 3 respectively. So for the $`\widehat{x}`$-oriented links we need 3 masks for each equal-$`x`$ hyper-plane (i.e., a three-volume here) and we have two independent equal-$`x`$ hyper-planes, giving a total of 6 masks for each Cartesian direction for the link orientation. Since there are 4 orientations, then there is a total of 24 masks needed for an action containing both $`1\times 1`$ and $`1\times 2`$ plaquettes. Thus a single lattice sweep must take at least 24 sequential serial calculations even on the most parallel computing architecture. The masking procedure outlined here for this action can only be implemented when the number of lattice points in each dimension is a multiple of three. Inspection of Fig. 8 reveals the periodicity of three is required to maintain separation of links at the boundary. Since simulations are usually carried out on lattices with even numbered sides, this restricts the length of the lattice sides to multiples of six. Fortunately, multiples of four are easily obtained as described in the next section. Moreover, Sec. VI reports a high-performance mask for this action with a periodicity of four. It is interesting to note that when implementing this masking procedure on the CM-5 we achieved optimum performance by calculating the updates for all links on the lattice and by then only implementing those updates that were appropriate for the particular mask being used at the time. In other words for the lattices that we have studied so far on the CM-5 it was more efficient to calculate link updates that were never used, than it was to split the masked links over the various processor nodes and update only these masked links. This was due to the fact that there was a large overhead of communication time in assigning the masked links across the processors. The point of this observation is that the optimal use of the masks will in general depend on the details of the parallel computing architecture being used. ## V Masking the Lattice When Using a Generalized Improved Action. We can now generalize the algorithm presented in Sec. IV for arbitrarily improved planar actions. Let us begin as before by considering the update of links oriented in the $`\widehat{x}`$-direction. Let us assume that we have an action with $`n\times m`$ links where the $`n`$ refers to the $`\widehat{x}`$ direction and the $`m`$ refers to the $`\widehat{y}`$, $`\widehat{z}`$, $`\widehat{t}`$ directions. We will eventually argue that only the $`n_{\mathrm{max}}\times n_{\mathrm{max}}`$ case, where $`n_{\mathrm{max}}`$ is the greater of $`n`$ and $`m`$, is necessary in the general case. As shown in Fig. 9 the nearest simultaneously updatable links are separated by $`n`$ steps in the $`\widehat{x}`$ direction and $`(m+1)`$ steps in the other three Cartesian directions. Hence we see that we can write in our notation for the four Cartesian orientations of the links that $`\widehat{x}:\widehat{x}n\widehat{x},\widehat{y}(m+1)\widehat{y},\widehat{z}(m+1)\widehat{z}\mathrm{and}\widehat{t}(m+1)\widehat{t},`$ (19) $`\widehat{y}:\widehat{x}(m+1)\widehat{x},\widehat{y}n\widehat{y},\widehat{z}(m+1)\widehat{z}\mathrm{and}\widehat{t}(m+1)\widehat{t},`$ (20) $`\widehat{z}:\widehat{x}(m+1)\widehat{x},\widehat{y}(m+1)\widehat{y},\widehat{z}n\widehat{z}\mathrm{and}\widehat{t}(m+1)\widehat{t},`$ (21) $`\widehat{t}:\widehat{x}(m+1)\widehat{x},\widehat{y}(m+1)\widehat{y},\widehat{z}(m+1)\widehat{z}\mathrm{and}\widehat{t}n\widehat{t}.`$ (22) We can now follow the arguments of the previous section. Let us consider a fixed-$`x`$ hyper-plane (i.e., three-volume). In place of a $`3\times 3`$ three-volume we will now need an $`(m+1)\times (m+1)\times (m+1)`$ three-volume. Furthermore, we will need a complete set of orthogonal and diagonal masks for this. Let us again look along the $`\widehat{x}`$ direction at a fixed $`t`$ plane for now, i.e., we are looking at a $`\widehat{y}`$$`\widehat{z}`$ plane as in Fig. 10. Let us refer to the $`(m+1)\times (m+1)`$ two-dimensional plane with the updatable links (solid dots) along the diagonal as plane 1. Then we can generate the other $`m`$ two-dimensional planes by diagonal half-shifts as before as depicted in Figs. 11 and 12. We can then sequentially stack these planes in the $`\widehat{t}`$ direction as before to form the first of the three-dimensional masks. The other $`m`$ three-dimensional masks are then generated from this first mask by the cyclic permutations of the $`m+1`$ planes as in Sec. IV. Hence we have generated the desired complete set of $`(m+1)`$ orthogonal three-dimensional diagonal masks. So for each fixed $`x`$-hyper-plane (i.e., three volume) we need $`(m+1)`$ masks. We will need such a set of masks for the $`n`$ values of $`x`$. The general result is that for updating the links oriented in the $`\widehat{x}`$ direction we need a total of $`n\times (m+1)`$ masks and we have seen that the construction of these masks is straightforward. The construction of the masks for the other Cartesian orientations of the links proceeds identically. This total number of masks is $`n_{\mathrm{mask}}=4\times n\times (m+1)`$. The periodicity of the mask is governed by the last factor, $`(m+1)`$, and the lengths of the lattice dimensions must be a multiple of this number. The reason for this is that if this were not the case then the imposition of the necessary periodic boundary conditions would cause link collisions, where a link being updated uses one or more other links which are simultaneously being updated. Any improved lattice action of physical interest must be both $`Z_4`$-symmetric (i.e., symmetric under the arbitrary interchange of the four Cartesian directions) and translationally invariant. Thus for such actions every link will find itself occurring in every possible position for every plaquette in the improved action. We then see, as we did in Sec. IV and Fig. 5, that the number of steps needed in each direction is determined by the longest plaquette side appearing in the action. Let us denote the longest plaquette side appearing in the action as $`n_{\mathrm{max}}`$. Then we see that the number of steps needed in the various Cartesian directions is given by $`\widehat{x}:\widehat{x}n_{\mathrm{max}}\widehat{x},\widehat{y}(n_{\mathrm{max}}+1)\widehat{y},\widehat{z}(n_{\mathrm{max}}+1)\widehat{z}\mathrm{and}\widehat{t}(n_{\mathrm{max}}+1)\widehat{t},`$ (23) $`\widehat{y}:\widehat{x}(n_{\mathrm{max}}+1)\widehat{x},\widehat{y}n_{\mathrm{max}}\widehat{y},\widehat{z}(n_{\mathrm{max}}+1)\widehat{z}\mathrm{and}\widehat{t}(n_{\mathrm{max}}+1)\widehat{t},`$ (24) $`\widehat{z}:\widehat{x}(n_{\mathrm{max}}+1)\widehat{x},\widehat{y}(n_{\mathrm{max}}+1)\widehat{y},\widehat{z}n_{\mathrm{max}}\widehat{z}\mathrm{and}\widehat{t}(n_{\mathrm{max}}+1)\widehat{t},`$ (25) $`\widehat{t}:\widehat{x}(n_{\mathrm{max}}+1)\widehat{x},\widehat{y}(n_{\mathrm{max}}+1)\widehat{y},\widehat{z}(n_{\mathrm{max}}+1)\widehat{z}\mathrm{and}\widehat{t}n_{\mathrm{max}}\widehat{t}.`$ (26) Hence the number of masks in general for an improved action will then be given by $$n_{\mathrm{mask}}=4\times n_{\mathrm{max}}\times (n_{\mathrm{max}}+1)$$ (27) and the lattice will need the length in each dimension to be an integral multiple of $`(n_{\mathrm{max}}+1)`$. It is useful to note that the linear masking for the standard Wilson action is the one that is extended initially in Sec. IV and is subsequently generalized in this section. For the standard Wilson action (i.e., $`1\times 1`$ plaquettes only) we see that $`n_{\mathrm{max}}=1`$ and hence $`n_{\mathrm{mask}}=4\times 1\times 2=8`$ as we found for the linear (and checkerboard) mask. For the improved action that we have studied (i.e., $`1\times 1`$ and $`1\times 2`$ plaquettes) we have $`n_{\mathrm{max}}=2`$ and hence $`n_{\mathrm{mask}}=4\times 2\times 3=24`$ or 6 masks per link direction as found in Sec. IV. However, this way of proceeding for the plaquette plus rectangle improved action would require each lattice dimension be a multiple of $`(n_{\mathrm{max}}+1)=3`$, but since we also typically want our lattices to have even lengths then that means each side of the lattice would need to be a multiple of 6 in length. Since the result in Eq. (27) is a lower bound, we can of course always choose to enlarge the period of our masking by choosing $`n_{\mathrm{max}}+2`$ for the last factor in Eq. 27 rather than $`n_{\mathrm{max}}+1`$. This will still ensure that no link collisions occur. For example, for the plaquette plus rectangle improved action we can use $`(n_{\mathrm{max}}+2)=4`$ instead of $`(n_{\mathrm{max}}+1)=3`$ in Eq. (27), so that any lattice lengths which are multiples of 4 become available at the cost of requiring 32 masks rather than 24. Fortunately, for this case a more efficient mask can be realized and will be presented in the next section. ## VI Non-planar Considerations We have presented a method for identifying links which may be simultaneously updated during Monte-Carlo updates or cooling sweeps. The generality of the algorithm allows one to parallelize link updates for planar actions of any degree of non locality. In this section we extend this analysis to a few special cases of actions in which out-of-plane considerations are necessary. Both cases are centred around the plaquette plus rectangle action of Eq. (2) in which $`1\times 1`$ and $`1\times 2`$ Wilson loops are considered in the action. Such actions dominate current improved gauge action analyses. In Sec. IV we illustrated how such an action can be masked through the consideration of an elementary $`3\times 3\times 3`$ cube in which one-third of the links may be simultaneously updated. However, only every second link in the direction of the links is updated simultaneously as illustrated in Fig. 6. Hence six masks per link direction are required. Here we consider an alternative masking specialized to the $`1\times 1`$ and $`1\times 2`$ Wilson loop actions. Fig. 13 illustrates the manner in which these Wilson loops may be nested, such that one need not restrict the mask to every second link in the direction of the links being updated. This technique will reduce the number of masks by a factor of two, at the expense of considering an elementary $`4\times 4\times 4`$ cube in which one-quarter of the links may be simultaneously updated. Fig. 14 displays the four planes to be cycled through in which the links to be updated simultaneously are indicated by the solid dot. Hence only four masks per link direction are required. Moreover, the lattice dimensions (usually even numbers) can now be multiples of four as opposed to six. The out-of-plane considerations required for the nested action are also indicated in Fig. 13. Hence it becomes apparent that not only the three links at $`(x,y+1)`$, $`(x,y+2)`$, and $`(x,y+3)`$, be avoided, but also the links two-steps in a direction orthogonal to the link direction and one step in a third direction (similar to moves of a Knight on a chess board) must be avoided. Inspection of the four planes to be cycled through in the elementary $`4\times 4\times 4`$ cube displayed in Fig. 14 indicates that such Knight moves are already avoided in this mask. However, it also becomes clear that the ordering of the planes is crucial. For example interchanging the positions of planes 2 and 3 would cause “link collisions” within the nested mask. Finally we consider non-planar actions in which one step out of the plane of the $`1\times 1`$ and $`1\times 2`$ Wilson loops is required. Such non-planar paths are introduced to eliminate small but finite $`𝒪(g^2a^2)`$ errors where $`g`$ is the gauge coupling constant. The six-link paths commonly referred to as the “chair” and “parallelogram” introduce a link parallel to that being updated which is one-step orthogonal to the link direction and one step in a third direction. Inspection of Fig. 14 indicates that such 1 by 1 moves eliminates fully two of the four planes and half of the parallel sites on each surviving plane. An example of four of the sites which may still be updated in parallel are indicated by the circled sites in Fig. 14. As a result there are now 16 masks required per link direction instead of 4. Now a total of 64 masks is required for this action which is still regarded as rather local. The introduction of even the most local non-planar paths can have a serious detrimental effect on the level of parallelism that is possible. It is easy to see that one can rapidly eliminate all sites in an elementary $`n\times n\times n`$ cube with non-planar loops, leading to $`n^3`$ masks per link direction. ## VII Summary and Conclusion We have briefly described the concept of improved actions and have explained the implications of the non locality arising from the improvement program for the implementation of these actions on parallel computing architectures. We have characterized these implications in terms of the number of masks, which in turn determine the minimum number of serial calculations needed to perform a Monte Carlo updating sweep over all of the gluon links on the lattice. We have systematically built up a completely general algorithm using masks that allow one to put any planar improved lattice action on a parallel machine in an efficient way. The generalized masking construction are given in Sec. V. Non-planar considerations encountered in nesting specific planar actions and actions involving non-planar loops have also been addressed. We hope that the methodology presented will allow one to find an efficient parallel mask for any desired action. We are currently testing our algorithms on some highly improved actions and will be reporting the results of these studies in the near future. ## VIII Acknowledgment. This research was supported by the Australian Research Council and by grants of supercomputer time on the CM-5 made available through the South Australian Centre for Parallel Computing. AGW also acknowledges support from the U.S. Department of Energy Contract No. DE-FG05-86ER40273 and by the Florida State University Supercomputer Computations Research Institute which is partially funded by the Department of Energy through Contract No. DE-FC05-85ER2500.
warning/0001/astro-ph0001179.html
ar5iv
text
# A deep ROSAT survey XV. The average QSO spectrum and its evolution. ## 1 Introduction Following the discovery of X-ray luminous QSO $`25`$ years ago (Lampton et al. 1972), we remain without a detailed description of the form and emission mechanism of QSO X-ray spectra, despite extensive studies of these objects. Satellite missions such as Ginga and EXOSAT indicated that the X-ray emission of radio-quiet QSO (which constitute the vast majority of the QSO population) above $`2`$ keV can be well described by a power-law with a photon index of $`\mathrm{\Gamma }1.9`$ (Williams et al. 1992, Lawson et al. 1992). However, the sources detected by these experiments which are the X-ray brightest, typically with $`f_\mathrm{x}1\times 10^{12}\mathrm{erg}\mathrm{cm}^2\mathrm{s}^1`$, nearby ($`z<0.2`$) QSO may not be representative of the class as a whole. Reeves et al. (1997) observed a small number of bright radio-quiet QSOs with ASCA, extending these results to higher redshifts. Recent long-exposure ($`100`$ ksec) ASCA observations (Georgantopoulos et al. 1997) extend the investigation of the average QSO X-ray spectra down to even fainter flux limits of $`5\times 10^{14}\mathrm{erg}\mathrm{cm}^2\mathrm{s}^1`$ in the $`210`$ keV band. Although, the photon statistics are not sufficient to be conclusive, there is some evidence that the spectral index of these faint QSOs is flat ($`\mathrm{\Gamma }=1.5\pm 0.2`$). Observations of the soft X-ray spectrum ($`<2`$ keV) of radio quiet, optically selected QSOs made with the Einstein observatory show a steeper spectral index ($`\mathrm{\Gamma }>2`$). This suggests that the QSO spectra are concave i.e. they cannot be represented by a single power-law over a broad energy band (eg Schwartz & Tucker 1988). This steep soft spectral index could be interpreted as excess emission (soft excess), at energies below $`1`$ keV, above the extrapolation of the X-ray power-law slope found at higher energies. Such a soft excess is detected in a substantial number of quasars (Masnou et al. 1992, Saxton et al. 1993), and in about 50 per cent of nearby Seyfert type AGN (Turner & Pounds 1989). Ciliegi & Macaccaro (1996) examine the X-ray spectral properties of 63 X-ray selected AGN, from the Einstein Medium Sensitivity Survey, finding $`\mathrm{\Gamma }2.4`$ with an intrinsic dispertion in spectral slope of $`\sigma 0.4`$, suggesting that X-ray spectral slopes are independent of whether selection takes place at either X-ray or optical wavelengths. After the launch of ROSAT (Trümper 1982), sensitive in the range $`0.12.4`$ keV and carrying instruments with improved energy resolution over previous missions, information on the soft X-ray properties of QSO burgeoned, due, in part, to large surveys with flux limits ($`5\times 10^{15}\mathrm{erg}\mathrm{cm}^2\mathrm{s}^1`$) orders of magnitude fainter than those of previous surveys. Schartel et al. (1996a) studied the stacked (average) spectra of a sample of 908 objects from the LBQS sample of optically selected QSOs. They find a mean spectral slope of $`\mathrm{\Gamma }2.5`$ and marginal evidence for a flattening of the spectral index at higher redshifts. Schartel et al. (1996b) analysed the individual spectra of a sample of 55 radio-quiet QSOs from the ROSAT all-sky survey. They find again steep spectral indices $`\mathrm{\Gamma }2.5`$ and no evidence for spectral evolution with redshift. Laor al. (1997) also derive a steep spectra index $`\mathrm{\Gamma }2.72\pm 0.09`$ using a sample of 19 radio-quiet, optically selected QSOs with $`z<0.4`$. The above values for the spectral slope of QSOs constitute the discrepancy known as the spectral paradox (Boldt 1987), in that the most common class of X-ray source has a spectral index significantly steeper than that of the X-ray background (XRB), which has $`\mathrm{\Gamma }1.5`$ at both hard (Marshall et al, 1980, Gendreau et al. 1995) and soft energies (Georgantopoulos et al. 1996), and so cannot be the major contributor to the XRB flux. A resolution of the spectral paradox could be the evolution of the mean QSO X-ray spectrum with redshift. Here, we derive the average QSO spectral index using a sample of X-ray selected QSOs detected in seven fields from our deep ROSAT survey (Shanks et al. 1991, Georgantopoulos et al. 1996). Previous work by Almaini et al. (1996) in five of our fields finds no variation with redshift in the slope of the spectra of QSOs (as determined by analysis of hardness ratios in the $`0.52.0`$ keV band) from five deep ROSAT fields. In order to identify the effect of a soft excess (emission significant only at $`\stackrel{<}{}0.5`$ keV) on the resultant power-law slope in the $`0.12.0`$ keV band, we use here the full ROSAT energy band and also perform proper spectral fits to the data. In this paper we include two further deep ROSAT fields in our investigation, creating a sample of 165 QSOs covering a broad redshift range, $`0.1z3.2`$, with a relatively high mean redshift ($`\overline{z}1.5`$). We also include two longer expsores (50 ksec each) of the QSF1 and QSF3 fields, which, along with the increased range of energies considered, results in photon statistics that are effectively doubled over those in the Almaini et al. work. Preliminary results from this work were presented in Stewart et al. (1994). The goals are to a) detect and quantify the possible evolution of the QSO spectral index b) compare the QSO spectra with that of the XRB c) examine the possibility for an upturn of the average spectral index in the soft band. ## 2 X-ray data selection We use data taken from deep observations of seven fields with the ROSAT PSPC (a summary of the observations taken is shown in table 1). Over 300 sources were detected within the inner-ring of the detector, at the $`4\sigma `$ level, down to a flux of $`4\times 10^{15}\mathrm{erg}\mathrm{cm}^2\mathrm{s}^1(0.52.0`$ keV). After an optical identification program with the AAT, 165 QSOs are identified, see Georgantopoulos et al. (1996) for the full details of the identification procedure and Shanks et al. (in preparation) for the catalogue of sources. Redshifts are in the range $`0.1z3.2`$. Fig. 1 shows the distribution. Datasets used in this work are available from the LEDAS ROSAT public database. We exclude data from periods of high particle background i.e. when the Master Veto Rate is above $`170\mathrm{count}\mathrm{s}^1`$ (Plucinsky et al. 1993), typically $`10`$ per cent of the data is rejected. Correction of the spectra for vignetting and the PSF were peformed with the Asterix software package. A background spectrum was accumulated from the region inside the inner-ring of the detector, with all sources detected above the $`4\sigma `$ level masked out and using a circular region with a radius of 90 arcsec. The search for spectral evolution of an average QSO requires binning the data into sufficient redshift subsamples to detect trends over the redshift range. However, one must strike a balance between the number of redshift bins and the resultant number of objects and photons in each bin. Clearly the number of photons must be adequate to allow for full spectral fitting, also, a low number of sources per bin may result in the spectrum being dominated by one particular source, rather then representing the spectrum of a mean source. Details of the redshift bins used, including the total number of background subtracted photons in each bin, are given in table 2. In addition to these redshift-segregated spectra, we created a spectrum including data from all 165 QSOs. We note also that the photons were distributed evenly over the fields, thus no one observation or field will dominate the results. One of the two observations of the QSF3 field, and one of the two observations of the QSF1 field was made with the PSPC-C instrument, so with these data we use the `pspcc_gain1_256.rsp` matrix. We use the `pspcb_gain2_256.rsp` matrix for all other observations, as they are made after October 1991 with the PSPC-B. Spectral fitting is performed for spectra from each field and instrument simultaneously. ## 3 Spectra and results Initially we considered the energies in the range $`0.52.0`$ keV, and fitted a simple power-law plus absorption column model to the data, for comparison with the results from Almaini et al. (1996). The Galactic $`N_\mathrm{H}`$ for all our fields (Stark et al. 1992, see table 1) is reasonably low, with a range of $`1.23.0\mathrm{cm}^2`$ and an exposure weighted mean of $`N_\mathrm{H}`$$`2.0\times 10^{20}\mathrm{cm}^2`$. Hereinafter, unless otherwise stated, for the Galactic $`N_\mathrm{H}`$ used in model fitting, we use appropriate $`N_\mathrm{H}`$ for each of the fields included in each of the spectra. The best-fit power-law indices found from the stacked photons from all redshift bins is $`\mathrm{\Gamma }2.23\pm 0.07`$ (table 3), with a $`\chi _\nu ^2`$ value of 1.07 from 157 degrees of freedom, in excellent agreement with Almaini et al. Errors quoted throughout this paper are 90 per cent errors calculated on the basis of the conservative assumption that all spectral parameters (including normalisation) are interesting. Following this we fitted the same model to data from each redshift bin individually, resulting in best-fit photon indices presented graphically in Fig. 2, and in tabular form in table 4. A $`\chi ^2`$ test shows that this model is formally acceptable in each redshift bin, furthermore, we find a non-evolving spectral slope, consistent, within errors, in each redshift bin, with the value for $`\mathrm{\Gamma }`$ found when considering data from all redshift bins together. These results are also consistent with the Almaini et al. results for the QSO spectral slope as a function of redshift \[see Fig. 3b of Almaini et al. (1996)\]. We then tested our data over the range $`0.12.0`$ keV with the same model. The derived photon indices are shown in the upper panel of Fig. 3, with the $`\chi _\nu ^2`$ given in the lower panel. For comparison, the best-fit photon index found for data taken from all QSOs is represented by the dashed line in the upper panel, with the $`1\sigma `$ error shown by the dotted lines. Evolution of the derived power-law photon index is found, in that $`\mathrm{\Gamma }_{(z<1.0)}>2.50`$ and $`\mathrm{\Gamma }_{(z>1.0)}2.25`$. To establish the significance of this difference in spectral slope, we performed spectral fits on two further spectra, including data from QSOs in the redshift range $`0.01.0`$ and $`1.03.2`$ respectively (for results see table 3). This difference is significant at over the $`8.5\sigma `$ level. It is, however, easily seen that this model provides an increasingly inadequate description of the data for smaller redshifts, indeed a $`\chi ^2`$ test rejects the model fit in the two lower redshift bins to at least the $`99.9`$ per cent level. The typical shape of a spectrum that rejects this model is shown in Fig. 4, the residuals (lower panel) clearly indicating a systematic difference between the data and the model at energies $`\stackrel{<}{}0.5`$ keV. To determine whether the data requires absorption in excess of the Galactic value, as seen in a small number of AGN (Ciliegi & Maccacaro 1996), we allowed the absorbing column density to be a free parameter in the fit, again performing the fit for both PSPC-B and PSPC-C spectra simultaneously, and with independent column densities. Fig. 5 shows how the best fitting column density varies with redshift for the PSPC-B spectra (uppermost panel) and the PSPC-C spectra (middle panel), with the dashed line representing the appropriate weighted mean Galactic $`N_\mathrm{H}`$ and the dotted lines the estimated uncertainty in the mean. The total absorbing column is consistent with the galactic value in all the redshift bins. Neither inclusion of $`N_\mathrm{H}`$ as a free parameter nor including an intrinsic warm-absorber reduce the $`\chi ^2`$ significantly, as determined by an F-test (Bevington & Robinson 1994), in any of the redshift bins. In an attempt to fit the $`0.12.0`$ keV spectra, and guided by results from nearby AGN which have a spectral upturn towards low energies (eg Turner and Pounds, 1989, Fiore et al. 1994), we tried a two-component model consisting of a power-law plus a redshifted black-body, the derived photon indices and $`\chi _\nu ^2`$ are shown in Fig. 6 and table 4. We see a significant reduction in the $`\chi ^2`$ due to the inclusion of this soft-excess even though the model is still formally rejected in the lowest redshift bin. However this result is probably not unreasonable, as we have co-added the photons from many sources at slightly different redshifts (and having different physical parameters and geometries), thus a single black-body model may not provide a perfect description of the data. We find that the power-law slope derived when including a black-body soft-excess is, in each redshift bin, consistent within the 90 per cent errors with $`\mathrm{\Gamma }1.8\pm 0.2`$, the value found for the entire sample of QSOs together. This value is consistent, within the 90 per cent error bars, with the intrinsic AGN power-law index of $`\mathrm{\Gamma }1.9`$ (Nandra & Pounds 1994), suggesting that the hard power-law continues down to below $`0.5`$ keV in the QSO rest frame. The best-fit temperatures of the black-body components (kT $`100200`$ eV) are in general agreement with those found for quasars (Rachen et al. 1996) and nearby radio-quiet QSOs (Fiore et al. 1994). Modelling the soft-excess as a thermal bremsstrahlung, a hot plasma, or an additional power-law yields similar $`\chi ^2`$ results over the five redshift bins, and derived power-law indices that, again, are around $`\mathrm{\Gamma }1.8`$. While we note that our description of the soft-excesses found in the binned spectra as a simple single-temperature black-body is simply an expedient to obtain a better statistical representation of the data and as such may not be fully representative of the true spectra of individual objects, or the underlying physical properties, it is of interest to examine the behaviour of the black-body parameters as a function of redshift bin. In Fig. 7 and table 5 we present the best fit black-body temperature and the luminosity of the black-body component as a function of redshift (now fixing the power-law photon index at 2 in order to better constrain the black-body properties). It is clear that the black-body temperature, while consistent with measurements for nearby objects of similar luminosity (Fiore et al. 1994) at zero redshift, increases in temperature to higher redshifts, rising from $`100`$ eV in the most local bin to of order 200 eV at redshifts $`2`$. In contrast, the mean luminosity of the black-body per object remains approximately constant. We first ask whether the effect could be some instrumental artefact, as the black-body temperature in the observer’s frame remains roughly constant. A systematic error in the calibration of the effective area, for example, could cause such an effect. We can rule out this possibility by noting that the fractional contribution from the soft excess relative to the power-law component varies with redshift from bin to bin, while a systematic uncertainty or error in the calibration would give a constant fractional contribution. Another possibility is that the results we obtain are an artefact of our assumption that the spectra of QSOs within a redshift bin can be represented by a single value of $`\mathrm{\Gamma }`$. For example coadding a step and a flat spctrum would result in a steep and flat spectrum at low energies and high energies respectively i.e. in a concave spectrum. In principle, deriving the spectra of individual QSOs would circumvent this problem. Fiore et al. (1994) find that a large fraction of the QSOs in their sample (four out of six) reveal a soft excess at low PSPC energies. In contrast Laor et al. (1997) find that the PSPC spectra of the majority of their LBQS QSOs are consistent with a single power-law to within 30%. Unfortunately, our objects typically are faint and we cannot derive individual spectra. However, as an exercise we have derived individually the spectra of the two brightest QSOs in the QSF3 field. The column density was fixed at the Galactic value $`N_H=1.7\times 10^{20}\mathrm{cm}^2`$. For RXJ0342.6-4404 (z=0.377) we obtain $`\mathrm{\Gamma }=3.18_{0.08}^{+0.08}`$ for a single power-law fit ($`\chi ^2=59.9/51`$ dof). The addition of a black-body component (0.06 keV) is significant at the 99 per cent confidence level ($`\chi ^2=51.0/49`$ dof) with $`\mathrm{\Gamma }=2.58_{0.21}^{+0.25}`$. In contrast RXJ0342.0-4403 (z=0.635) can be fit with a single power-law ($`\mathrm{\Gamma }=2.9_{0.10}^{+0.20}`$) without the need of an additional component ($`\chi ^2=40.3/38`$ dof). It is evident that a conclusive answer on the origin of the soft excess cannot be given with the photon statistics curently available. Finally, it is possible that a range of individual QSO spectra could result in artificially induced behaviour which varies with redshift due to sampling different energy ranges in each redhsift bin. If for example QSOs had a range of spectral indices then sampling at different energy ranges would bias the results in that using higher energy bands flatter spectra would be preferentially be sampled giving rise to an apparent flattening of spectrum with redshift. Such an explanation for the behaviour observed here seems contrived and unlikely given the lack of variation in spectra in bins with redshifts greater than $`1`$ . Next, we consider the possible physical origin of the spectral upturn. The soft excesses observed in AGN spectra are often ascribed to the high energy end of the big blue bump seen in the optical/UV region of the spectrum and thought to originate in the accretion disk (Walter & Fink 1993, Turner & Pounds 1989). The properties of the accretion disk are then dominated by the mass accretion rate and mass of the central black-hole. In the simplest models of the emission from the accretion disk, the spectrum is a multi-temperature black-body with the maximum temperature component coming from the inner disk radii. The behaviour of the soft-excess parameters presented here certainly suggest that the emission is more complex than would be expected in a thin accretion disk. The high temperatures found do not fit well with the above models. Moreover, the constant luminosity of the soft excess seen in the ROSAT band while the power law component increases by a factor of 20-30 is more consistent with up-scattering than with thermal emission from a disk powered by potential energy loss. We note that the constant luminosity of our measured soft excesses is similar to that found by Saxton et al. (1993) for a sample of nearby QSOs. They found that the soft excess luminosity saturated above a power law luminosity of $`10^{44}\mathrm{erg}\mathrm{s}^1`$, the mean luminosity in our lowest redshift bin. They postulated that this was due to a decreasing temperature of the soft-excess. Such an explanation is contradicted by our results, which are more in keeping with an increasing mass-accretion rate rather than black-hole mass being the important factor. Regardless of the origin of the spectral upturn at low energies, the average QSO spectrum has important implications for the origin of the XRB. Boldt et al. (1987) argue that as the AGN have hard (2-10 keV) X-ray spectra, significantly steeper than the spectrum of the XRB in the same band (Marshall et al. 1980), they cannot produce the bulk of the XRB. This spectral paradox appears to extend at the lower ROSAT energies. The XRB in the 0.5-2 keV band also has this flat spectral index, $`\mathrm{\Gamma }1.5`$, (Georgantopoulos et al. 1996, Vecchi et al. 1999). All previous studies of the QSO spectra in the ROSAT band (eg Schartel et al. 1996, Laor et al. 1997) obtain spectral indices much steeper than the XRB spectrum in the same band. However, these contain mainly bright, low redshift QSOs. In contrast, our sample contains typical QSOs in the sense that our objects constitute a large fraction (over $`50`$ per cent at 1 keV) of the XRB and they cover a wide range of redshift and luminosity. Our results clearly show that the soft X-ray selected QSOs cannot produce the bulk of the soft XRB. Another population with a flat X-ray spectrum is needed at faint fluxes. This population may be associated with AGN obscured in X-ray wavelenghts; some examples of this population at low redshift ($`z<0.5`$) may have been already identified in deep ROSAT pointings (Schmidt et al. 1998). ## 4 Conclusions We derived the average QSO spectrum and its evolution with redshift, using a sample of 165 QSOs from 7 deep ROSAT fields ($`f_\mathrm{x}>4\times 10^{14}\mathrm{erg}\mathrm{cm}^2\mathrm{s}^1`$). We find strong evidence ($`>8\sigma `$) for spectral evolution, in that the $`0.12.0`$ keV spectra flatten from $`\mathrm{\Gamma }2.6`$ at low redshifts, to $`\mathrm{\Gamma }2.1`$ at $`z2.4`$. At high redshifts a power-law model with Galactic absorption describes the stacked spectra well, with the power-law index in rough agreement with the slope found from the $`0.52.0`$ keV spectra. However, for z$`<1.5`$, the spectra of the QSOs in the $`0.12.0`$ keV band cannot be fitted by a single power-law, with Galactic absorption, in contrast with some other ROSAT results on low redshift QSOs (eg Laor et al. 1997) Furthermore, we find that these deviations from a simple power-law cannot be modelled by excess $`N_\mathrm{H}`$ absoption along the line of sight. The above suggest the presence of a soft excess below 0.5 keV in the QSO rest-frame. Indeed, when we add a black-body component to model the soft excess, we obtain a significant reduction in the $`\chi ^2`$ for the low redshift bins. However this model is still rejected in the lowest redshift bin, probably because a single temperature black-body cannot represent adequately the coadded soft excess of all QSOs in one bin. The resulting power-law component has a photon index $`\mathrm{\Gamma }1.8\pm 0.2`$, and shows no evidence of evolution with redshift. This spectral index is not inconsistent, within the 90 per cent errors, with both the canonical AGN spectral index of $`\mathrm{\Gamma }1.7`$ observed in the $`210`$ keV band, and the intrinsic spectral slope of $`\mathrm{\Gamma }1.9`$, suggesting that the hard X-ray power-law may continue well into the ROSAT band, down to below $`0.5`$ keV. The soft excess probably evolves with redshift, maintaining a constant luminosity but increasing in temperature from 100 eV locally, to 200 eV at $`z2`$. We caution that the black-body model is simply an expedient to obtain a better statistical representation of the data. It is possible that the soft excess observed may be just an artefact of the coaddition of spectra with a large range of spectral indices. Regardless of the origin of the soft excess, the average QSO spectrum has important implications for the origin of the XRB. It is shown here that the average spectrum of the faint, high redshift QSOs is much steeper than the spectrum of the XRB in the soft 0.1-2 keV band, extending the spectral paradox to low energies. Future missions such as XMM are expected to bring a breakthrough in the spectral studies of high redshift, faint QSOs due to the broad energy coverage and high effective area. ## 5 acknowledgments AJB acknowledges the receipt of a University of Leicester studentship. IG and OA acknowledge the support of PPARC. ### REFERENCES Almaini, O., Shanks, T., Boyle, B. J., Griffiths, R. E., Roche, N., Stewart, G. C., Georgantopoulos, I., 1996, MNRAS, 282, 295 Bevington, P. R., Robinson, D. K., 1994, Data reduction and error analysis for the physical sciences. $`\mathrm{M}^\mathrm{c}`$Graw-Hill International Editions, New York Boldt, E., 1987, Phys. Rep., 146, 215 Boyle, B. J., Griffiths, R. E., Shanks, T., Stewart, G. C., Georgantopoulos, I., 1993, MNRAS, 260, 49 Ciliegi, P., Maccacaro, T., 1996, MNRAS, 282, 477 Fiore F., Elvis M., $`\mathrm{M}^\mathrm{c}`$Dowell J. C., Siemiginowska A., Wilkes B. J., 1994, ApJ, 431, 515 Gendreau, K., et al., 1995, PASJ, 47, L5 Georgantopoulos, I., Stewart, G. C., Shanks, T., Griffiths, R. E., Boyle, B. J., 1993, MNRAS, 262 619 Georgantopoulos, I., Stewart, G. C., Shanks, T., Boyle, B. J., Griffiths, R. E., 1996, MNRAS, 280, 276 Georgantopoulos, I., Stewart, G.C., Blair, A.J., Shanks, T., Griffiths, R.E., Boyle, B.J., Almaini, O., Roche, N., 1997, MNRAS, 297, 203 Lampton M., Margon, B., Bower, S., Mahoney, W., Anderson, K., 1972, ApJL, 171, 45 Laor, A., Fiore, F., Elvis, M., Wilkes, B. J., Mcdowell, J. C., 1997, ApJ, 477, 93 Lawson, A. J., Turner, M. J. L., Williams, O. R., Stewart, G. C., Saxton, R. D., 1992, MNRAS, 259, 743 Marshall, F. E., et al., 1980, ApJ, 235, 4 Masnou, J. L., Wilkes, B. J., Elvis, M., Mcdowell, J. C., Arnaud, K. A., 1992, å, 253, 35 Matt, G., Fabian, A. C., Ross, R. R., 1993, MNRAS, 264, 839 Nandra, K., Pounds, K. A., 1994, MNRAS, 268, 405 Plucinsky, P. P., Snowden, S. L., Briel, U. G., Hasinger, G., Pfeffermann, E., 1993, ApJ, 418, 519 Rachen J. P., Mannheim K. & Biermann P. L., 1996. å, 310, 371 Raymond J. C., Smith B. W., 1977, ApJS, 35, 419 Reeves, J.N., Turner, M.J.L., Ohashi, T., Kii, T., 1997, MNRAS, 292, 468 Reynolds, C. S., 1997, MNRAS, 286, 513 Ross, R. R., Fabian, A. C., Mineshige, S., 1992, MNRAS, 258, 189 Saxton, R. D., Turner, M. J. L., Williams, O. R., Stewart, G. C., Ohashi, T., Kii, T., 1993, MNRAS, 262, 63 Schartel, N., Green, P.J., Anderson, S.F., Hewett, P.C., Foltz, C.B., Margon, B., Brinkmann, W., Fink, H., Trumper, J., 1996a, MNRAS, 283 1015 Schartel, N., Walter, R., Fink, H. H., Truemper, J., 1996b, å, 307, 33 Schmidt, M. et al., 1998, A&A, 329, 495 Schwartz, D.A., & Tucker, W.H., 1988, ApJ, 332, 157 Shanks T., Georgantopoulos I., Stewart G. C., Pounds K. A., Boyle B. J., Griffiths R. E., 1991, Nat, 353, 315 Stark A. A., Gammie C. F., Wilson R. W., Bally J., Linke R. A., Heiles C., Hurwitz M., 1992, ApJS, 79, 77 Trümper, J., 1982, Adv. Space Res., 2, 241 Turner, T. J., Pounds, K. A., 1989, MNRAS, 240, 833 Vecchi, A., Molendi, S., Guainazzi. M., Fiore, F., Parmar, A.N., 1999, A&A, 349, 73 Walter, R., Fink, H. H., 1993, å, 274, 105 Wilkes, B. J., Elvis, M., 1987, ApJ, 323, 243 Williams, O. R., et al., 1992, ApJ, 389, 157
warning/0001/hep-th0001072.html
ar5iv
text
# 1 Introduction ## 1 Introduction The standard action for a particle or string on a coset space $`G/H`$ is manifestly invariant under $`G`$ but does not have a quadratic kinetic term. This obstructs the usual quantization procedure. Moreover, the isometries are nonlinearly realized on the coordinates and so even if the action were quadratic, the fields would not automatically form $`G`$-representations. This makes it difficult to directly study systems such as strings on $`AdS_{p+2}\times S^{dp2}`$ which are important for understanding holography. A hint on how to work around this comes from twistors. These were originally set up by Penrose to study conformal Minkowski space, and have since been generalized to conformal superspace and $`AdS_5`$. In all of these cases, twistors associated a hypersurface in some vector space which transforms linearly under the isometry group with every point of the coset space. The internal coordinates of these hypersurfaces were associated with momenta and constrained quantities. This construction can be generalized to arbitrary coset manifolds<sup>1</sup><sup>1</sup>1The results below apply both to cosets and supercosets, with no additional restrictions (reductivity, symmetry, etc.) except where explicitly noted. $`G/H`$. A mapping between points of the coset and hypersurfaces in a vector space can be constructed which naturally mimics the geometric structure of the coset. The isometries, for example, can easily be extracted from the linear isometry transformations of the twistors. If the vector space is also a Hilbert space (i.e. posesses an appropriate inner product) then one can naturally construct objects out of twistors which are manifestly invariant under the coset isometries. Using the twistor mapping, these can be written as (typically fairly complicated) functions of the coset coordinates and the internal coordinates of the twistor. Since these quantities are manifestly $`G`$-invariant, one can construct actions out of them which are equivalent to ordinary coset actions if the internal coordinates are identified with momenta. Since the twistor mapping is typically very complicated, very simple twistor actions are equivalent to very complicated coset actions. We demonstrate this construction for a particle on $`AdS_p`$. Twistors are built in a vector space which transforms in the spinor representation of $`SO(p1,2)`$. A world-line gauge theory can be built out of these twistors which is equivalent to the ordinary action for a massive particle on the coset. This theory is equivalent to a nonlocal $`\sigma `$-model whose target space is the vector space. This construction can probably be generalized to the study of particles and strings on anti-de Sitter superspaces such as those important for the AdS/CFT correspondence. ## 2 The twistor construction We begin by describing cosets in a language which naturally leads to twistors. A point in a coset $`G/H`$ is associated with a hypersurface in the group manifold by the relation $$\varphi (\widehat{x}G)=\{\widehat{x}h:hH\}.$$ (1) The $`\widehat{x}`$ which generate distinct $`\varphi (\widehat{x})`$ are given by $$\widehat{x}:=v(\xi )\widehat{x}_0,$$ (2) where $`\widehat{x}_0G`$ is the origin of the coset space, (an arbitrary point) and $`\xi `$ is a collective coordinate on $`G/H`$. The function $`v(\xi )`$ is a coset representative, which for our purposes is a function from the coordinates to the group such that $`v(0)=1`$ and $`\varphi v`$ is 1-1, so distinct hypersurfaces are associated with distinct coordinates. A particular form of $`v(\xi )`$ which we will often use is $$v(\xi )=e^{\xi K}h(\xi ),$$ (3) where the $`K`$ are the generators of $`G`$ not in $`H`$, and $`h(\xi )`$ is some $`H`$-valued function chosen to simplify the resulting expression. Using this function, we can associate coset coordinates with hypersurfaces in the group manifold in a natural way: $$\varphi (\xi )=\varphi (\widehat{x}(\xi ))=v(\xi )\varphi (\widehat{x}_0).$$ (4) These hypersurfaces are naturally invariant under the right action of $`H`$. Such a construction cannot always be globally performed. The problem is analogous to the selection of coordinates on a sphere $`S^2=SO(3)/SO(2)`$. Technically, it arises because the coset representative $`v(x)`$ is a section of the principal bundle $`GG/H`$ and so in general cannot be globally defined. We may resolve this issue analogously to the problem of coordinates by covering $`G/H`$ with patches (open sets whose intersections are contractible) and performing this construction on each open set. Transition functions on intersections are naturally induced from the transition functions of the principal $`G`$-bundle of which $`v`$ is a section. It is also necessary to choose a different $`\widehat{x}_0`$ for each patch, which is analogous to using the north and south poles as origins of the two coordinate systems on $`S^2`$. The result of such a construction will (as we will see below) be a twistor bundle rather than a global twistor space, but this will not introduce any unfamiliar complexities. Using this construction, the geometry on the coset space may be defined by the invariances of the Cartan form $`L=v^1dv`$. This form is canonically separated into $`L=EK+\mathrm{\Omega }H`$, where $`E`$ is the vielbein and $`\mathrm{\Omega }`$ the $`H`$-connection. (This generalizes the spin connection of Minkowski space) When $`G`$ is semisimple, this can be contracted with a restricted Cartan-Killing metric to give a metric on the coset; $$g^{\mu \nu }=\eta ^{AB}L_A^\mu L_B^\nu ,$$ (5) where the indices $`A`$ and $`B`$ run over only the coset ($`K`$) generators of $`G`$.<sup>2</sup><sup>2</sup>2This procedure is discussed in detail in . In the more general case (which includes Minkowski space) the procedure is somewhat more subtle. For some groups at least, there is an invariant symmetric two-form which may take the place of the Cartan-Killing metric $`\eta `$; however, there is no general existence proof nor is there a method of computing such forms. In this case one assumes that the transformations which lead to covariant transformations of the $`K`$-components of the Cartan form will become isometries of the coset spaces. The isometries of this space are given implicitly by the action of $`G`$ on the coset representative: $$\delta \xi :\delta v(\xi )=gv(\xi ),$$ (6) where $`gG`$. This implies that $$\delta \varphi (\xi )=g\varphi (\xi ).$$ (7) It is straightforward to compute the actual transformations of the $`\xi `$ from this relationship if we write (7) in an explicit representation. This motivates us to define twistors to be explicit group representations of these hypersurfaces $`\varphi (\xi )`$. Specifically, if we represent the group on a Hilbert space $`\mathrm{\Lambda }`$, (not necessarily infinite-dimensional!) a twistor is a mapping of coordinates to $`H`$-invariant hypersurfaces in $`\mathrm{\Lambda }`$ given by $$𝒵(\xi )=v(\xi )𝒵_0,$$ (8) where $`v(\xi )`$ is the $`\mathrm{\Lambda }`$-representation of the coset representative, and $`𝒵_0`$ is an $`H`$-invariant hypersurface in $`\mathrm{\Lambda }`$. (The $`\mathrm{\Lambda }`$-representation of $`\varphi (\widehat{x}_0)`$) As before, the mapping $`𝒵`$ must be 1-1 for the set of $`𝒵(\xi )`$ to be isomorphic to the coset, which means that the codimension of $`𝒵_0`$ in $`\mathrm{\Lambda }`$ must be no less than the dimension of the coset. ($`\mathrm{codim}𝒵(\xi )=\mathrm{codim}𝒵_0`$ for any $`\xi `$ since $`v(x)`$ is surjective) We will also restrict ourselves to $`\mathrm{dim}𝒵_0>0`$, since otherwise $`𝒵`$ would be a mapping of points onto points and so would lose several of the interesting features which we will discover below. We next write $`𝒵_0`$ explicitly as a linear function of some coordinates on $`\mathrm{\Lambda }`$. (These coordinates may be curvilinear, although we do not consider such possibilities in depth here) Then using the explicit form of the coset representative $`v(\xi )`$, we may write each $`𝒵(\xi )`$ as a function of the coordinates $`\xi `$ and the internal coordinates $`\lambda `$ of $`𝒵_0`$ which is linear in $`\lambda `$ and typically fairly complicated in $`\xi `$. This process has two advantages. First, since the $`𝒵`$ are given as explicit functions of the coset coordinates, it is straightforward to use (6) to compute the geometric properties of the space. This is especially valuable in the case of complicated cosets such as $`AdS_p\times S^{dp}`$ superspace, where traditional (differential-equation) methods of calculating isometries are very difficult. Second, since $`\mathrm{\Lambda }`$ is a Hilbert space there is a natural continuous and complete inner product of twistors which is manifestly invariant under the action of $`G`$. This allows one to easily construct quantities with a very complicated dependence on $`\xi `$ and $`\lambda `$ which are invariant under $`G`$. This invariance persists even though $`HG`$ is typically nonlinearly realized on the coset space. If (as we will do later) we identify the $`\lambda `$ with some internal parameters of a system such as momenta, it is possible to use these invariants to construct very simple twistor-based actions which are equivalent to very complicated coordinate-space actions. As the preceding discussion was somewhat abstract, it is useful to consider some explicit examples. We begin with the case of conformal Minkowski space $`SO(3,2)/ISO(3,1)\times D`$, where $`D`$ is the dilatation operator. We choose as our representation the 4-component spinor representation of $`SO(3,2)`$, which decomposes into a 2-component spinor and a 2-component conjugate spinor of $`ISO(3,1)`$ of conformal weights $`\pm 1/2`$.<sup>3</sup><sup>3</sup>3Although we use spinorial representations here, this is by no means a general feature of twistors. In this representation, a group element has the form $$g=\left(\begin{array}{cc}L_\beta {}_{}{}^{\alpha }+\frac{1}{2}D\delta _\beta ^\alpha & iK_{\alpha \dot{\alpha }}\\ iP^{\dot{\alpha }\alpha }& \overline{L}^{\dot{\alpha }}{}_{\dot{\beta }}{}^{}\frac{1}{2}\delta ^{\dot{\alpha }}_{\dot{\beta }}\end{array}\right)$$ (9) and the initial hypersurface $`𝒵_0`$ has the form $$𝒵_0=\left(\begin{array}{c}\lambda _\alpha \\ \mu ^{\dot{\alpha }}\end{array}\right),$$ (10) where the $`\alpha `$ ($`\dot{\alpha }`$) are (conjugate) spinor indices of $`SO(3,1)`$ and $`\lambda `$ and $`\mu `$ are complex. The stability group $`H`$ is generated by the $`L`$, $`K`$, and $`D`$. Lorentz invariance implies that if $`𝒵_0`$ has any point with $`\lambda 0`$, it must contain all such points, and likewise for $`\mu `$. Thus the dimension constraint $`0<\mathrm{dim}𝒵_04`$ requires that exactly one of the two be independent. Without loss of generality, we choose $`\lambda `$ and let $`\mu `$ be a linear function thereof. The remaining part of $`H`$-invariance then determines $`\mu =0`$. Now let us choose the canonical coset representative $$v(x)=e^{ixP}=\left(\begin{array}{cc}1& 0\\ ix^{\dot{\alpha }\alpha }& 1\end{array}\right).$$ (11) The twistor mapping is now $$𝒵(x)=\left(\begin{array}{c}\lambda _\alpha \\ ix^{\dot{\alpha }\alpha }\lambda _\alpha \end{array}\right).$$ (12) This is the familiar Penrose twistor formula. We will not discuss isometries and invariants in this case, saving that instead for the more detailed example of $`AdS_p`$ below. It is worth noting that this procedure was by no means unique. The freedoms of choice are in the selection of an appropriate coset representative (which will typically be determined by algebraic simplicity, subject to the requirement that $`𝒵(\xi )`$ is 1-1) and in the choice of the initial hypersurface $`𝒵_0`$. We can also naturally ask about the invariants which may be constructed out of these twistors. The simplest world-line action which one may construct out of these twistors is clearly $$=i\overline{𝒵}𝒵,$$ (13) where contraction has been performed with the standard spinor metric. If we substitute in (12), and write $$P_{\dot{\alpha }\alpha }=\overline{\lambda }_{\dot{\alpha }}\lambda _\alpha ,$$ (14) then this action reduces to the simple form $$S=i𝑑\tau Px$$ (15) which is the usual world-line action for a massless particle. $`P`$ is automatically null in this case because the spinor metric $`ϵ^{\alpha \beta }`$ is antisymmetric. Massive actions cannot easily be written in terms of these twistors, which is unsurprising since we are here working in conformal Minkowski space. In this case we have put the internal coordinates $`\lambda _\alpha `$ of the twistor to use as momenta of the particle. It is not clear how general such an interpretation is; clearly a precondition for the possibility of so doing is that the twistor bundle (the set of these hypersurfaces $`𝒵(x)`$ over every point, with open sets as discussed above) contains the tangent bundle of $`G/H`$ as a subbundle. Even when this is not possible, the procedure above will turn the $`\lambda _\alpha `$ into Lagrange multipliers for various quantities; when the quantities are derivatives of the coordinates, there is a somewhat natural momentum interpretation. We will see more of this construction later. ## 3 Twistorization of $`AdS_p`$ We now turn to the case of particles on $`AdS_p=SO(p1,2)/SO(p1,1)`$. The ordinary world-line action for these particles is manifestly $`G`$-invariant but does not have a quadratic kinetic term, so it is useful to try to rephrase this in terms of twistors. This is reasonable since the first-order action, $$=\frac{1}{2}Px+P_\rho \rho +u\left[\frac{1}{2\rho ^2}P^2\rho ^2P_\rho ^2m^2R^2\right]$$ (16) contains only terms of the form $`Px`$, which are similar to those found in the conformal Minkowski action (13), and a constraint term which is $`G`$-invariant although not manifestly so. In a twistor construction one hopes that this can be rewritten in a manifestly symmetric (and preferably simple) way, and we will see that this is indeed the case. Twistorization must begin with a choice of $`G`$-representation. The two simplest choices are the fundamental and the spinor. The fundamental has simpler group generators, but since its dimension is $`(p+1)`$ such twistors would have only one internal coordinate and so momenta could not be encoded by the twistor. Therefore we use the spinor representation, which has complex dimension $`2^{(p+1)/2}2d`$. The group elements in this representation are<sup>4</sup><sup>4</sup>4Our index notation is: $`\mu =0\mathrm{}p2`$ is an $`SO(p2,1)`$ (Lorentz) vector index, $`\alpha ,\beta `$ and $`\dot{\alpha },\dot{\beta }=1\mathrm{}d`$ are Lorentz spinor and conjugate spinor indices, respectively. $`A,B`$ and $`\dot{A},\dot{B}=1\mathrm{}2d`$ are spinor and conjugate spinor indices of $`SO(p1,2)`$. $$g^A{}_{B}{}^{}=\left(\begin{array}{cc}L_\beta {}_{}{}^{\alpha }+\frac{1}{2}D\delta _\beta ^\alpha & iK_{\alpha \dot{\alpha }}\\ iP^{\dot{\alpha }\alpha }& \overline{L}^{\dot{\alpha }}{}_{\dot{\beta }}{}^{}\frac{1}{2}D\delta ^{\dot{\alpha }}_{\dot{\beta }}\end{array}\right)$$ (17) The $`K`$ and $`P`$ generate conformal transformations and conformal momentum, which are related to $`AdS`$ conformal transformations and momenta by $`\stackrel{~}{K}`$ $`=`$ $`(KP)/2`$ $`\stackrel{~}{P}`$ $`=`$ $`(K+P)/2.`$ (18) The $`L_\beta ^\alpha `$ generate the Lorentz group and the $`D`$ are dilatations. The stability group $`H`$ is generated by the $`L`$’s and the $`\stackrel{~}{K}`$’s. First we must choose an $`H`$-invariant initial hypersurface. We can write this surface in the form $$𝒵_0^A=\left(\begin{array}{c}\lambda _{0\alpha }\\ \mu _0^{\dot{\alpha }}\end{array}\right)$$ (19) As in the Penrose case, $`L`$-invariance requires that if $`𝒵_0`$ contains any independent points with $`\lambda _00`$, then it contains all such points, and similarly for $`\mu _0`$. Since we want $`0<\mathrm{dim}𝒵_02dp`$, only one of these two should be independent of the other. Without loss of generality we choose $`\lambda _0`$ to be independent, and fix $`\mu _0`$ by $$\mu _0^{\dot{\alpha }}=F^{\dot{\alpha }\beta }\lambda _{0\beta }+G^{\dot{\alpha }\dot{\beta }}\overline{\lambda }_{0\dot{\beta }}$$ (20) for some $`F^{\dot{\alpha }\beta }`$ and $`G^{\dot{\alpha }\dot{\beta }}`$ which parametrize our twistorization. $`\stackrel{~}{K}`$-invariance then requires that $`F\gamma _\mu F+G\gamma _\mu \overline{G}`$ $`=`$ $`\gamma _\mu `$ (21) $`F\gamma _\mu G+G\gamma _\mu \overline{F}`$ $`=`$ $`0`$ (22) where the $`\gamma _\mu ^{\alpha \dot{\alpha }}`$ are the Dirac matrices for $`SO(p2,1)`$. For simplicity we will consider the case $`F=0`$, so $$𝒵_0^A=\left(\begin{array}{c}\lambda _{0\alpha }\\ G^{\dot{\alpha }\dot{\beta }}\overline{\lambda }_{0\dot{\beta }}\end{array}\right).$$ (23) A simple choice of coset representative is $$v(x^\mu ,\rho )=e^{xP}e^{D\mathrm{log}\rho }=\left(\begin{array}{cc}\rho ^{1/2}& 0\\ i\rho ^{1/2}x^{\dot{\alpha }\alpha }& \rho ^{1/2}\end{array}\right);$$ (24) using this, and defining $`\lambda =\rho ^{1/2}\lambda _0`$, the twistor is $$𝒵^A(x^\mu ,\rho )=\left(\begin{array}{c}\lambda _\alpha \\ ix^{\dot{\alpha }\alpha }\lambda _\alpha +\rho ^1G^{\dot{\alpha }\dot{\beta }}\overline{\lambda }_{\dot{\beta }}\end{array}\right).$$ (25) As a check, the isometries of the space can be calculated from $$\delta 𝒵^A=g^A{}_{B}{}^{}𝒵_{}^{B}.$$ (26) Varying both sides of (25), one finds $$\delta \lambda _\alpha =\left(L_\alpha {}_{}{}^{\beta }+\frac{1}{2}D\delta _\alpha {}_{}{}^{\beta }+K_{\alpha \dot{\alpha }}x^{\dot{\alpha }\beta }\right)\lambda _\beta i\rho ^1K_{\alpha \dot{\alpha }}G^{\dot{\alpha }\dot{\beta }}\overline{\lambda }_{\dot{\beta }}$$ (27) and so $`\delta x^{\dot{\alpha }\alpha }`$ $`=`$ $`P^{\dot{\alpha }\alpha }x^{\dot{\alpha }\beta }L_\beta {}_{}{}^{\alpha }\overline{L}^{\dot{\alpha }}{}_{\dot{\beta }}{}^{}x_{}^{\dot{\beta }\alpha }+Dx^{\dot{\alpha }\alpha }+x^{\dot{\alpha }\beta }K_{\beta \dot{\beta }}x^{\dot{\beta }\alpha }\rho ^2K^{\dot{\alpha }\alpha }`$ (28) $`\delta \rho `$ $`=`$ $`D\rho 2\rho xK`$ (29) which are the well-known isometries of anti-de Sitter space. Geometric invariants may now be constructed by contracting $`𝒵`$ with the $`SO(p2,1)`$ metric $$H_{\dot{A}}{}_{}{}^{B}=\left(\begin{array}{cc}0& 𝒞^{\dot{\alpha }\dot{\beta }}\\ \overline{𝒞}^{\dot{\alpha }\dot{\beta }}& 0\end{array}\right)$$ (30) where $`𝒞`$ is the charge conjugation matrix, so $$\overline{𝒵}_1𝒵_2=\overline{\lambda }_1\mu _2+\overline{\mu }_1\lambda _2.$$ (31) A natural first guess for a particle action is $$=i\overline{𝒵}𝒵.$$ (32) This matches the kinetic term in (16) if we identify components of $`\lambda `$ with the momenta as follows: $`P_{\alpha \dot{\alpha }}`$ $`=`$ $`2\lambda _\alpha \overline{\lambda }_{\dot{\alpha }}`$ $`P_\rho `$ $`=`$ $`{\displaystyle \frac{i}{2\rho ^2}}\left[\overline{\lambda }G\overline{\lambda }\lambda \overline{G}\lambda \right].`$ (33) In (32), however, all the components of $`\lambda `$ are independent and so their dynamics must be specified. The first condition is the mass-shell constraint, $$\frac{1}{2\rho ^2}P^2\rho ^2P_\rho ^2=\frac{1}{4}\left(\overline{𝒵}𝒵\right)^2=M^2R^2.$$ (34) There exist further independent components of $`\lambda `$ for most values of $`p`$. These may be fixed by fixing the values of a set of twistor bilinears $$\varphi _i\overline{𝒵}T_i𝒵$$ (35) where $`(T_i)_{\dot{A}}^B`$ are some constant matrices which transform in the $`(\frac{1}{2},\frac{1}{2})`$ of $`SO(p1,2)`$. The number of independent $`\varphi _i`$ that must be set depends on $`p`$. In an action, these will be constrained to values $`m_i`$. For example, using (34) the mass-shell constraint is $`\varphi _{T=1}=2MR`$. So the complete twistor action takes the simple form $$=i\overline{𝒵}\left(iu^iT_i\right)𝒵u^im_i$$ (36) where the $`u^i`$ are Lagrange multipliers. This action is equivalent to (16). It has several important features: 1. The action is manifestly $`SO(p1,2)`$ invariant and has a quadratic kinetic term. It has the structure of a world-line gauge theory with sources. The “gauge fields” $`u^i`$ are nondynamical since there is no field strength in one dimension. This statement can be made somewhat more precise by noting that (32) implies that the Poisson brackets (which will become commutators in the quantized theory) are $$\{𝒵_A,\overline{𝒵}^{\dot{B}}\}_{PB}=2iH_A^{\dot{B}}$$ (37) with all other brackets vanishing, and so $$\{\varphi _i,\varphi _j\}_{PB}=2i\overline{𝒵}[T_i,T_j]𝒵.$$ (38) Since the set of constraints under Poisson brackets forms a Lie algebra, the set of $`T_i`$ form one as well, and this algebra is invariant under $`SO(p1,2)`$. This guarantees that the action (36) indeed has a gauge symmetry. 2. The gauge group contains a $`U(1)`$ factor corresponding to the mass-shell constraint $`T=1`$, $`m=2MR`$. The rest of the group may be calculated explicitly for small $`p`$ by constructing the $`\varphi _i`$; they are | p | 1 | 2 | 3 | 4 | 5 | 6 | 7 | | --- | --- | --- | --- | --- | --- | --- | --- | | $`\mathrm{dim}𝒵_0`$ | 2 | 2 | 4 | 4 | 8 | 8 | 16 | | $`N_\varphi `$ | 1 | 0 | 1 | 0 | 3 | 2 | 9 | | Group | $`U(1)^2`$ | $`U(1)`$ | $`U(1)^2`$ | $`U(1)`$ | $`U(1)\times SU(2)`$ | $`U(1)^3`$? | $`U(1)\times SU(2)^3`$? | The final two are conjectured but have not been explicitly calculated. This is related to the result of for $`AdS_5`$. In that case, the 8-component spinors were decomposed into a pair of 4-component spinors of the stability group $`H=SO(4,1)`$ indexed by $`I,J=1,2`$, and $$(T_i)_{aI}{}_{}{}^{bJ}=(\sigma _i)_I{}_{}{}^{J}𝒞_{a}^{}^b$$ (39) (The $`a,b`$ are $`SO(4,1)`$ spinor indices) 3. This twistor Lagrangian can be quantized following a procedure similar to that used in , leading to solutions which transform in representations of $`SO(p1,2)`$. 4. For $`i0`$, The $`\varphi _i`$ may be chosen to be independent of the momenta. In these cases it is not clear what meaning one could assign to a nonzero $`m_i`$. The analogous quantities in are all zero. 5. The Lagrange multipliers $`u^i`$ can be integrated out to give $$^{}(k)=i\overline{𝒵}(+iTm)𝒵|_k+\frac{dq}{2\pi }\left(\overline{𝒵}T^i𝒵\right)|_{k+q}\left(\overline{𝒵}T_i𝒵\right)|_{kq}$$ (40) which is therefore equivalent to (36). (This can also be seen by explicitly resumming Feynman diagrams involving the $`u^i`$) The actions (36) and (40) represent a considerable simplification over their classical counterpart (16). Because they have leading quadratic terms and manifest $`G`$-symmetry, their quantum solutions automatically fill out representations of the isometry group. A similar construction can be carried out for an arbitrary coset manifold, or even a supercoset, and (similarly to ) can be used to construct string actions on these spaces. Since the known superstring actions are manifestly invariant under the isometries, it is likely that these systems will be amenable to a twistor interpretation which would allow their quantization and analysis, including interactions with Ramond-Ramond and Neveu-Schwarz background fields. Acknowledgements The author wishes to thank Piet Claus, Renata Kallosh, Michael Peskin, Joachim Rahmfeld, and Steve Shenker for numerous useful discussions and comments, and the organizers of the TASI-1999 school where part of this work was accomplished. This work was supported in part by an NSF Graduate Research Fellowship and by NSF grant PHY-9870115.
warning/0001/hep-ex0001033.html
ar5iv
text
# Some problems in plasma suppression of beam-beam interactions at muon colliders. Talk at the Workshop Studies on Colliders and Collider Physics at the Highest Energies: Muon Colliders at 10 TeV to 100 TeV, 27 September - 1 October, 1999 Montauk, New York, USA, be published by the American Institute of Physics. ## 1 Introduction One of main the problems for high energy muon colliders is the limitation of the luminosity due to beam-beam interactions. A measure of the beam-beam interaction is the tune-shift parameter $`\xi `$ . For round beams, $`\xi =Nr_c/4\pi ϵ_n`$, where $`r_c=e^2/mc^2`$ is the classical radius of the beam particles, $`ϵ_n`$ is the normalized transverse emittance. The value of $`\xi `$ should be small enough ($`\xi _{max}<0.1`$) - otherwise the beams are disrupted due to resonance diffusion. The maximum luminosity $`L_{max}=N\gamma f\xi _{max}/r_c\beta `$, where $`\beta `$ is the $`\beta `$-function (usually $`\beta \sigma _z`$), $`f`$ is the collision rate. So, $`L\xi `$. This effect puts a severe limit on the luminosity of muon colliders. One of the possible solution of this problem is plasma suppression of beam-beam interactions . In a sufficiently dense plasma one can expect that the induced charges and currents will decrease the beam fields and, consequently, the beam-beam effects. If plasma decreases the beam field by a factor K, one can use beams with K times smaller $`ϵ_n`$ (to keep $`\xi =\xi _{max}`$), correspondingly the luminosity will be K times larger. It is essential to reduce both electric and magnetic fields because in the vacuum their action on the opposing beam are equal in value and direction, the effective beam field is $`|E|+|B|`$. If the plasma density $`n_p`$ is larger than the particle density of the colliding bunches $`n_b`$, the electric field of the beam will be suppressed by repelling (in the case of negatively charged bunch) or attracting (in the case of positively charge bunch) plasma electrons (ions are immobile). The nature of the magnetic field suppression is somewhat more complicated. In the linear approximation the resulting suppression of the beam-beam interaction $$\xi /\xi _04/(k_p\sigma _r)^2\text{for}k_p\sigma _r>>1,$$ (1) where $`k_p=\omega _p/c`$, $`\omega _p^2=4\pi n_pe^2/m_e`$, $`\sigma _r`$ is the r.m.s. beam radius. A more accurate result, including nonlinear effects and finite plasma thickness, was obtained in ref. . So, for a factor of 5 suppression of beam-beam interactions the plasma should satisfy the following requirements: a) $`n_p>n_b`$ and b) $`k_p\sigma _r>4`$. For example, let us take parameters of the “evolutionary” 100 TeV muon collider (see the B.King’s table) but with 5 times smaller $`ϵ_{nx}`$: $`N=0.8\times 10^{12}`$, $`\sigma _r=\sqrt{2}\sigma _x=0.13\mu \text{m},`$ $`\sigma _z=0.25`$ cm. In this case $`\xi =0.5`$ without suppresion, while the acceptable $`\xi =0.1`$. The plasma density required for decreasing $`\xi `$ by a factor of 5 is found from conditions a) and b), which give $`n_p>2.3\times 10^{21}`$ and $`n_p>2.5\times 10^{22}`$, respectively. As a source of plasma one can use a liquid Li jet with electron density $`1.5\times 10^{23}`$ cm<sup>-3</sup>. Such a target will be fully ionized by the muon beam and the return current. If this theory is correct, in the considered example one can increase the luminosity by a factor of 5. With other beam parameters one can expect even better results, up to 30 for the given beam diameter. All this sounds nice. However, there are two effects which create serious problems for this method, and, perhaps, close it: * collisions in the plasma; * hadronic background at large angles. ## 2 Collisions in the plasma. In all papers on plasma suppression of the beam-beam interaction it was assumed that plasma is collisionless. This picture is not correct. In fully ionized plasma the electrons of the return current do not lose energy on ionization, do not lose energy in collisions with other electrons, because all electrons move with the same average velocity, and have very small energy loss in collisions with the ions; however, their longitudinal velocity is decreased due to the scattering on the ions. The change of the longitudinal velocity in one scattering on the ion $$\mathrm{\Delta }v=2v_0\mathrm{sin}^2\frac{\vartheta }{2}v_0\frac{\vartheta ^2}{2}$$ (2) The resulting friction in the plasma $$\frac{d\stackrel{}{v}}{dt}=v_0\mathrm{\Delta }vn_p𝑑\sigma =v_0^2\frac{\vartheta ^2}{2}8\pi n_p\left(\frac{e^2Z}{m_ev_0^2}\right)^2\frac{d\vartheta }{\vartheta ^3}=$$ $$=4\pi n_p\frac{\stackrel{}{v}}{v_0}\left(\frac{e^2Z}{m_ev_0}\right)^2\mathrm{ln}\frac{\vartheta _{max}}{\vartheta _{min}}=4\pi n_p\frac{\stackrel{}{v}}{v_0}\left(\frac{e^2Z}{m_ev_0}\right)^2\mathrm{ln}\mathrm{\Lambda },$$ (3) where $`\mathrm{ln}\mathrm{\Lambda }=\mathrm{ln}(b_{max}/b_{min})`$. The minimum value of the impact parameter $`b`$ follows from the energy conservation $`b_{min}=e^2Z/mv_0^2`$, and the maximum value of $`b`$ is equal to the Debey length $`\lambda _D=\sqrt{kT/m}/\omega _p`$. For the considered plasma densities and electron velocities, $`\mathrm{ln}\mathrm{\Lambda }710`$. For estimation of the collision time one can take $`dv=v`$, which gives $$\tau _{col}\frac{v_0}{4\pi n_p\left(\frac{e^2Z}{m_ev_0}\right)^2\mathrm{ln}\mathrm{\Lambda }}.$$ (4) The average velocity of electron in the return current $`u(n_b/n_p)c`$. Although this velocity is not the same as $`v_0`$ due to transverse motion, let us the first assume $`v_0=u`$. Then $$\tau _{col}(\frac{n_b}{n_p})^3\frac{1}{4\pi cn_pZ^2r_e^2\mathrm{ln}\mathrm{\Lambda }}.$$ (5) For the example given above $`n_b/n_p1/60`$, $`Z=3`$, $`n_p=1.5\times 10^{23}`$, that gives $`\tau _{col}10^{17}`$ sec, which is much smaller than the bunch collision time $`\sigma _z/c0.25/3\times 10^{10}10^{11}`$ sec. So, the assumption of collisionless plasma is not valid. The plasma should be considered as a medium with some conductivity $`\sigma _c`$. Accurate calculation of conductivity is a complicated task because the electron drift in the longitudinal induction electric field with very small energy loss, only scatter. Due to the field their total kinetic energy continuosly grows, while the drift velocity is approximately constant: $`uc(n_b/n_p)`$. Nevertheless, we can make some estimate. Eq.3 is approximately valid even in the case of the “hot” return current if we will consider $`\stackrel{}{v}`$ as the drift velocity (from now on $`u`$) and $`v_0`$ as the “thermal” velocity, which is still unknown. The loss of the drift velocity given by Eq.3 is compensated by the induction electric field $$|du/dt|=eE_{||}/m_e.$$ (6) The conductivity is defined by equation $$en_pu=\sigma _cE.$$ (7) Drift velocity is known: $$u=c(n_b/n_p).$$ (8) Using Eqs 6,7,8, we get $$\sigma _c=\frac{e^2n_bc}{m_e|du/dt|}$$ (9) Using Eq.3, we obtain $$\sigma _c=\frac{m_ev_0^3}{4\pi e^2Z^2\mathrm{ln}\mathrm{\Lambda }}$$ (10) Now we have to estimate $`v_0`$. The collision time is given by eq.4. Between two collisions the electron drift velocity is restored by the induction electric field and the total energy is increased by about $`m_eu^2`$, so as an estimate one can take the average kinetic energy to be equal to $`m_ev_0^2N_{col}m_eu^2`$. The maximum number of scatterings for the same electron can be estimated as the number of collisions after which its transverse displacement is equal to the beam radius (then this electron in the return current is replaced by the new one which comes from outside and is initially cool) $$\tau _{col}v_0\sqrt{N}\sigma _r.$$ (11) Using this arguments and Eq.10 we find $$v_0=c(4\pi n_bZ^2\sigma _rr_e^2\mathrm{ln}\mathrm{\Lambda })^{1/5}$$ (12) For Z=3 (Li), $`n_b=2.5\times 10^{21}`$ (see the example above) and $`\sigma _r=0.15`$ $`\mu `$m, $`v_00.075`$c. So, the conductivity is found, see Eqs 10,12. Now we have to understand how the conductivity influences the plasma suppression of the bunch field. The return current is driven by the longitudinal electric field $`E_{||}`$ that is caused by penetration of the beam magnetic field into the plasma. From Faraday law $$E_{||}\frac{1}{c}\frac{d(B_\varphi \sigma _r)}{dt}.$$ (13) This electric field produces the return current equal approximately to the bunch current $`I`$ (if compensation works) $$\sigma _cE_{||}\pi \sigma _r^2I.$$ (14) Introdusing $`B_\varphi 2I/c\sigma _r`$ (the beam field when there is no beam field suppression) and Eqs 13,14, we obtain $$\frac{dB_\varphi }{B_\varphi }\frac{c^2dt}{2\pi \sigma _r^2\sigma _c}.$$ (15) The relative value of the beam field which penetrates into the plasma during the time of the bunch collision is $$\frac{\mathrm{\Delta }B}{B}\frac{c\sigma _z}{2\pi \sigma _r^2\sigma _c}\frac{2\sigma _zr_eZ^2\mathrm{ln}\mathrm{\Lambda }}{\sigma _r^2(v_0/c)^3},$$ (16) where $`v_0`$ is given by Eq.12. For the example considered in this paper: $`\sigma _z=0.25`$ cm, $`Z=3`$, $`\sigma _r=0.13`$ $`\mu `$m, $`N=0.8\times 10^{12}`$, $`v_0/c0.075`$, $`\mathrm{ln}\mathrm{\Lambda }=7`$ we get $$\frac{\mathrm{\Delta }B}{B}100!!!$$ (17) In order to obtain plasma suppression by one order of magnitude we need $`\mathrm{\Delta }B/B0.1`$. So, it seems that plasma does not help. Although my estimate is very approximate, it is very unlikely that a factor of 1000 is lost. ## 3 Backgrounds ### 3.1 Photo-nuclear reactions The total photo-nuclear cross section for lithium is $$\sigma _{\gamma Li}0.4\times 10^{27}\text{cm}^2.$$ (18) The number of virtual photon with the energy above 1 GeV per one muon at 100 TeV muon collider is $$N_\gamma \frac{2\alpha }{\pi }\mathrm{ln}\left(\frac{E}{\omega }\right)\frac{d\omega }{\omega }0.3N_\mu .$$ (19) The number of ph.n. reactions per bunch crossing generated by $`210^{12}`$ muons (two beams) in $`l=0.5`$ cm Li jet is $$N_b=N_\gamma n_{Li}l\sigma _\gamma 0.3\times 210^{12}\times 510^{22}\times 0.5\times 0.410^{27}=0.610^7!$$ (20) Although most of the produced particles travel in the forward direction, each reaction produce gives approximately one particle ($`\pi ^\pm ,\pi ^0`$) at large angles, with $`PP_t300`$ MeV. The total energy of these particles is greater than $`210^3`$ TeV. It is hard to imagine a detector which could work in conditions so terrible! ### 3.2 e<sup>+</sup>e<sup>-</sup> production: $`\mu Li\mu Li\text{e}\text{+}\text{e}\text{-}`$ The cross section of this reaction $$\sigma \frac{28\alpha ^2r_e^2}{27\pi }(l^36.36l^2)(Z_1Z_2)^2,$$ (21) where $`l=\mathrm{ln}\frac{2(P_1P_2)}{m_1m_2}\mathrm{ln}2\gamma _\mu `$. In the case of 50 TeV muons and the Li target, $`\sigma =1.8\times 10^{26}`$ cm<sup>2</sup> The probability of e<sup>+</sup>e<sup>-</sup> pair creation by a muon in a 1 cm thick Li jet for 1000 crossings (as it is in the muon colliders) is about $`1e^1`$ (one interaction length). In most cases the energy loss is not large but sufficient to knock the muon out of the $`10^4`$ energy range which contributes to luminosity (muons with larger energy deviations are defocussed due to chromatic abberations of the final focus system). So, this effect will decrease the luminosity lifetime by about a factor of 2. ## 4 Conclusions Suppression of the beam-beam effects by a dense plasma jet (Li) at the collision point is a very attractive idea. However, collisions in the plasma significantly change the picture. This effect was ignored before. Rough estimates show that this effect leads to fast “recovery” of the magnetic beam field, leaving it practically unsuppressed. This result should be checked by more accurate calculations. Photo-nuclear reactions produce enormous hadronic backgrounds in the detector ($`10^7`$ particles/crossing at large angles), so the possibility of experimentation at such background conditions is practically impossible. Electro-production of e<sup>+</sup>e<sup>-</sup> pairs in Li jet leads to some decrease of the luminosity lifetime. I would like to thank K.Lotov and A.Skrinsky for useful discussions and B.King for organization of the very fruitful workshop.
warning/0001/hep-th0001015.html
ar5iv
text
# 1 Introduction ## 1 Introduction In this talk the first steps in a rigorous study of topological quantum theories using canonical methods is described. A topological theory is characterised by a high degree of symmetry, with the number of symmetries equal to the number of fields of the theory. As a result the system has no dynamical degrees of freedom, and the space of fields modulo symmetries, which is trivial at the linear level, is typically a finite-dimensional moduli space which encodes ‘topological’ information. As with quantum theories in general, a powerful tool for studying a topological theory is functional integration. The generic expression for the vacuum generating functional takes the form $$_{\mathrm{fields}/\mathrm{symmetries}}𝒟\varphi \mathrm{exp}\left(iS[\varphi ]\right)$$ (1) where $`S(\varphi )`$ is the action of the field $`\varphi `$. Using standard (non-rigorous) methods of quantum field theory a number of new and unexpected mathematical results have been been derived from topological models, results which in many cases have then been fully proved by more standard mathematical methods, but which would probably not have been discovered without the insights gained from the quantum field theory. (An early appearance of topological invariants in the quantum field theoretic situation is due to Belavin, Polyakov, Schwarz and Tyupin . A more recent example of the powerful application of topological quantum field theory in mathematics may be found in , while fuller accounts of earlier work in this field may be found in the books of Nash and Schwarz .) Most functional integrals such as (1), and related expressions with operator insertions, have not at present been properly defined. However, since these integrals have such astonishing mathematical power, it seems that an attempt to define these objects rigorously should be more than worth while. In this talk we show how this may be done for the simplest topological model, the topological particle, and describe briefly some recent work by Hrabak which might lead to progress in the canonical quantization of topological field theories. Some rigorous results on path integrals (that is, functional integrals in quantum mechanics) are known. The basic classical result (which is described by Simon in ) for a particle of unit mass moving in one dimension with Hamiltonian $$H=\frac{1}{2}p^2+V(x)$$ (2) gives the action of the imaginary time evolution operator $`\mathrm{exp}(Ht)`$ on a wave function $`\psi (x)`$ by the formula $$\mathrm{exp}(Ht)\psi (x)=d\mu \mathrm{exp}(_0^tV((x(s))ds)\psi (x(t))$$ (3) where $`d\mu `$ denotes Wiener measure starting from $`x`$, and $`x(t)`$ are corresponding Brownian paths; the potential $`V`$ must satisfy certain analytic conditions. The curved space analogue of this result for a Riemannian manifold has been developed by Elworthy and by Ikeda and Watanabe . The expression for evolution according to the Hamiltonian $`H=L+V(x)`$ where $`L`$ is the scalar Laplacian looks identical to (3), but with $`x(t)`$ a process depending on metric and connection rather than simply flat space Brownian motion. Tangent space geometry plays an essential part in the theory. The present author has further extended these methods by developing a flat space theory of fermionic path integrals and marrying it with Brownian motion on manifolds to give Brownian motion on supermanifolds in a suitable form for handling the Hodge-de Rham operator and the Dirac operator on manifolds . ## 2 The Topological particle Following Beaulieu and Singer we consider the quantum-mechanical model with fields which are maps $`x:IM`$, where $`I`$ is the interval $`[0,t]`$ and $`M`$ is an $`n`$-dimensional Riemannian manifold with metric $`g`$. The action of the theory is $$S[x(.)]=_0^ti\omega _\mu (x(t^{}))\dot{x}^\mu (t^{})dt^{}$$ (4) $`\omega =dh`$ is an exact one form on $`M`$ with local co-ordinate expression $`\omega =\omega _\mu (x)dx^\mu `$ and $`\dot{x}^\mu (t^{})=\frac{dx^\mu }{dt^{}}`$. The action can be expressed in the simpler form $`S[x(.)]=i(h(x(t))h(x(0)))`$ which shows that the action is indeed highly symmetric, being independent of all but the endpoints of the paths $`x(t)`$. While Beaulieu and Singer consider the case where $`\omega =0`$, we consider the case where $`h`$ is a Morse function on $`M`$ (so that $`\omega =dh`$ is only zero at isolated points). To carry out the canonical quantization we first evaluate the momentum $`p_\mu `$ conjugate to $`x^\mu `$, obtaining $$p_\mu =\frac{\delta }{\delta \dot{x}^\mu }=i\omega _\mu ,$$ (5) which shows that the theory has $`n`$ constraints $$T_\mu p_\mu i\omega _\mu .$$ (6) The Hamiltonian of the theory is as usual defined to be $`H(p,x)=p_\mu x^\mu (x,\dot{x})`$, so that, as is generally the case for a topological theory, the Hamiltonian of the theory (prior to gauge-fixing) is zero. Since $`\omega `$ is closed these constraints are first class and abelian, that is $`\{T_\mu ,T_\nu \}=0`$ and $`\{T_\mu ,H_c\}=0`$. The infinitesimal gauge transformations generated by $`T_\mu `$ are $$\delta _ϵ\psi (x)=iϵ(_\mu \psi (x)+\omega _\mu (x)\psi (x))$$ (7) where $`\psi (x)`$ is a wave function and quantization is in the Schrödinger picture with $`p_\mu =i_\mu `$. (Below, when ghosts are introduced, we will find that we require $`p_\mu `$ to be represented as a covariant derivative $`i_\mu `$.) The explicit form of the gauge transformations suggests that representative of gauge equivalence classes may be obtained from the condition $`X^\mu \psi =0`$ where $`X^\mu =g^{\mu \nu }(p_\nu +i\omega _\nu )`$. The validity of these gauge-fixing conditions will become clear below. The BRST quantization scheme will now be applied to this system; to do this anticommuting ghosts $`\eta ^\mu `$ together with their conjugate momenta $`\pi _\mu `$ are introduced. (The phase space is now a $`(2n,2n)`$-dimensional supermanifold, with odd coordinates $`\eta ^\mu ,\pi _\mu `$ transforming as indices suggest.) Poisson brackets on this extended phase space are defined by the symplectic form $$dp_\mu dx^\mu +\pi _\mu \eta ^\mu +\frac{1}{2}dx^\mu dx^\nu R_{\mu \nu \kappa }{}_{}{}^{\lambda }\eta _{}^{\kappa }\pi _\lambda ,$$ (8) (where $``$ denotes covariant differentiation using the Levi-Civita connection); this is a special case of the symplectic form introduced by Rothstein . Quantization is carried out by introducing states represented by wave functions $`\psi (x,\eta )`$ and momenta acting as $`p_\mu =i_\mu `$, $`\pi _\mu =i\frac{}{\eta ^\mu }`$. The wave functions $`\psi (x,\eta )`$ are functions on the $`(n,n)`$-dimensional supermanifold $`SM`$ with local coordinates $`x^\mu ,\eta ^\mu `$, and the explicit form of the action of the covariant derivative on a wave function is given by $`_\mu \psi (x,\eta )=_\mu \psi (x,\eta )+\mathrm{\Gamma }_{\mu \nu }^\lambda \eta ^\nu {\displaystyle \frac{}{\eta ^\lambda }}\psi (x,\eta ).`$ (9) The BRST charge $`Q`$ takes the standard form $`Q=\eta ^\mu T_\mu =i\eta ^\mu (_\mu +\omega )`$. (A covariant derivative is not required here because of the symmetry of the connection.) The gauge-fixing fermion $`\chi `$ also takes the standard form $`\chi =\pi _\mu X^\mu =ig^{\mu \nu }\pi _\mu (_\nu \omega _\nu )`$. Alternatively, if we make the natural identification of forms on $`M`$ with wave functions $`\psi (x,\eta )`$ then $`Q=ie^hde^h`$, $`\chi =e^h\delta e^h`$ where $`d`$ denotes exterior differentiation of forms and $`\delta =d`$ is the adjoint operator. The gauge-fixing Hamiltonian is then $`H_g`$ $`=`$ $`i(Q\chi +\chi Q)`$ (10) $`=`$ $`d\delta +\delta d+g^{\mu \nu }\omega _\mu \omega _\nu i(\pi _\mu \eta ^\nu \eta ^\nu \pi _\mu ){\displaystyle \frac{^2h}{x^\mu x_\nu }}.`$ This Hamiltonian has appeared in the literature on other occasions; for instance, when $`h`$ is constant, it is the Hamiltonian used by Alvarez-Gaumé to prove the Atiyah-Singer index theorem. (A rigorous version of this proof may be found in .) It is also the first supersymmetric Hamiltonian used by Witten in his study of Morse theory . It is evident that the choice $`\chi =\pi _\mu X^\mu `$ is a good gauge-fixing condition, satisfying the essential conditions derived by the author in . First, as observed by Beaulieu and Singer in the constant $`h`$ case, the standard theory of harmonic forms shows that the gauge condition determines a unique element of each $`Q`$ cohomology class. (The observation that these arguments extend to all functions $`h`$ is due to Witten .) Also, the zeros of $`H_g`$ coincide with these representatives of the cohomology classes, while the eigenvalues of $`H_g`$ tend to infinity, so that this Hamiltonian does regulate the non-physical states. The path integral formulae for this Hamiltonian can be put in rigorous form using the methods of the author in , and used to establish rigorous results for this model. The key idea in this approach is to use Brownian paths $`x_t,\eta _t`$ in the supermanifold $`SM`$ with local coordinates $`x^\mu ,\eta ^\mu `$. The Brownian paths are defined by the stochastic differential equations $`x_t^\mu `$ $`=`$ $`x^\mu +{\displaystyle _0^t}e_{a,s}^\mu 𝑑b_s^a,`$ $`e_{a,t}^\mu `$ $`=`$ $`e_a^\mu +{\displaystyle _0^t}e_{a,s}^\nu e_{b,s}^\lambda \mathrm{\Gamma }_{\nu \lambda }^\mu (x_s)db_s^b`$ $`\eta _t^\mu `$ $`=`$ $`\eta ^\mu +\theta _t^ae_{a,t}^\mu `$ $`+`$ $`{\displaystyle _0^t}(\eta _s^\nu \mathrm{\Gamma }_{\nu \lambda }^\mu e_{b,s}^\lambda db_s^b\theta _t^ade_{a,s}^\mu +{\displaystyle \frac{1}{4}}\eta _s^\nu R_{\nu \lambda \kappa }{}_{}{}^{\mu }(x_s)\eta _s^\lambda \rho _s^ae_{a,s}^\kappa ds),`$ (11) where $`b_t`$ is flat bosonic Brownian motion and $`(\theta _t,\rho _t)`$ is flat fermionic Brownian motion. The measure corresponding to this process incorporates as the ‘kinetic term’ the heat kernel of the Laplace-Beltrami operator $`(d+\delta )^2`$, so that these Brownian paths are appropriate for the analysis of the Hamiltonian (10). Further details will be found in , where it will be shown that Witten’s approach to Morse theory can be put on an entirely rigorous mathematical footing. (Some parts of Witten’s analysis have been proved rigorously by Simon et al and by Mathai and Wu ; however the explicit modeling of the manifold’s cohomology via critical points and instanton calculations does not appear to have received a full mathematical treatment.) ## 3 The two-dimensional topological sigma model To conclude, a brief indication of some developments in a two-dimensional topological model will be described. The model, which was first proposed by Witten , concerns the geometry of $`J`$-holomorphic curves (or pseudoholomorphic maps) $`u:\mathrm{\Sigma }M`$ from a Riemann surface $`\mathrm{\Sigma }`$ into an a $`2m`$-dimensional almost-Kaehler manifold $`M`$. The other fields of the theory are a bosonic set $`H_\mu ^\alpha `$ and two fermionic sets $`\eta ^\mu `$ and $`\pi _\mu ^\alpha `$. (Here $`\alpha =1,2`$ are indices on $`\mathrm{\Sigma }`$ while $`\mu =1,\mathrm{},2m`$ are indices on $`M`$.) The fields $`H`$ and $`\pi `$ satisfy constraints $`P^{^{}}H=0,P^{^{}}\pi =0`$ where $`P^{^{}}{}_{\beta \nu }{}^{\alpha \mu }=\delta _\beta ^\alpha \delta _\nu ^\mu ϵ_\beta ^\alpha J_\nu ^\mu `$ is a projection operator with $`ϵ`$ the complex structure on $`\mathrm{\Sigma }`$ and $`J`$ the almost complex structures on $`M`$. Witten constructs by hand a set of supersymmetry transformations (beginning with $`\delta u^\mu =iϵ\eta ^\mu `$) on these fields, and then an invariant action. The model is used to derive deep geometric insights into the moduli space of $`J`$-holomorphic curves on $`M`$. Recent work of Hrabak shows that the rather complicated and seemingly ad hoc supersymmetry transformations of the model can be derived in the canonical setting as BRST transformations; the novel feature of Hrabak’s work is that the formalism used is not the standard canonical formalism (in which time plays a special rôle) but the multisymplectic formalism which is manifestly covariant; corresponding to the fields $`u:\mathrm{\Sigma }M`$ there are multimomenta $`p_\mu ^\alpha `$ (which after projection relate to Witten’s $`H_\mu ^\alpha `$), while in the ghost sector the ghosts have momenta $`𝒫_\mu ^\alpha `$ which relate to Witten’s $`\pi _\mu ^\alpha `$. The BRST symmetry obtained by Hrabak corresponds directly to the $`J`$-holomorphicity of the embeddings. Recent work by Kanatchikov on quantization in the multisymplectic framework suggests that it may be possible to use Hrabak’s approach to carry out a full canonical quantization of Witten’s interesting two-dimensional topological model.
warning/0001/nlin0001052.html
ar5iv
text
# 1 Introduction ## 1 Introduction Supersymmetric extensions of integrable hierarchies of PDE’s are by now a well-studied subject (see, e.g., ). In particular, for the KP equations, two different extensions have emerged: the Manin–Radul SKP (MRSKP), and the Jacobian SKP of Mulase and Rabin (JSKP). The most remarkable and best know differences between the two hierarchies are the fact that the flows of MRSKP are a representation of a super Heisenberg algebra, while those of JSKP are supercommuting ones, and the behavior of their algebro–geometric solution. Indeed, the JSKP describes linear flows on the (super) Jacobian manifold of a super algebraic curve $`\widehat{C}`$, while the MRSKP flows involve, in general, a motion on the space of moduli of $`\widehat{C}`$. Actually, the latter is perhaps the strongest motivation that led Mulase and Rabin to modify the (previously discovered) Manin–Radul supersymmetric extension of the KP hierarchy. From the point of view of the present paper, it is more relevant another difference between the two approaches: the Manin–Radul theory concerns the supersymmetric extension of the Lax representation: $$\frac{L}{t_k}=[\left(L^k\right)_+,L]$$ (1.1) of the KP hierarchy on the space of pseudo-differential operators in one dimension. In the approach of Mulase and Rabin, one starts instead from the Sato representation of the KP equations on the Volterra group of Dressing operators, $$\frac{S}{t_k}=\left(S_x^kS^1\right)_{}S,$$ (1.2) where $`S`$ and $`L`$ are related by the well–know dressing formula $$L=S_xS^1.$$ (1.3) This paper is based on the representation of the KP theory as a set of conservation laws , which has been recently studied as an outgrowth of the application of the Gel’fand–Zakharevich theorem to infinite dimensional integrable systems. Namely, the KP equations are written as the conservation laws $$\frac{h}{t_k}=_xH^{(k)},$$ (1.4) where $$h=z+\underset{l=1}{\overset{\mathrm{}}{}}\frac{h_l}{z^l}$$ (1.5) is the generating series of the conserved densities of the KP theory, and $`H^{(k)}`$ are their current densities. These currents are written as suitable linear combinations of the Faà di Bruno monomials $$h^{(k)}=\left(_x+h\right)^k1$$ (1.6) associated with $`h`$. The supersymmetric extension of the equations (1.4) will be written as $$\frac{\widehat{h}}{t_\alpha }=(1)^\alpha \delta \widehat{H}^{(\alpha )},$$ (1.7) and called Hamiltonian Super KP hierarchy (HSKP). Here $`\delta =_\phi +\phi _x`$ is the square root on the super-circle $`S^{1|1}`$ of the $`x`$–derivative $`_x`$, well known from string and superconformal field theories (see, e.g., ), the odd superfield $`\widehat{h}`$ replaces the generator $`h`$ of the local hamiltonian densities, and $`\widehat{H}^{(\alpha )}`$ are suitable supersymmetric extensions of the currents $`H^{(j)}`$. In the first part of the paper we show that, as it happens in the ordinary case, equations (1.7) admit a representation–extension to a family of dynamical systems with $`^2\times ^2`$ variables (of which half are even and half are odd), to be called the Super Central System (SCS). It is a counterpart of the Sato system on the infinite–dimensional Grassmann manifold , and the Super KP equations (1.7) can be considered as a kind of “reduction” of SCS. Then we show how the hierarchy (1.7) can be identified, by means of a non trivial coordinate change, with the Jacobian Super KP hierarchy. This is however only the first topic we want to discuss in this paper. The second one concerns the relation of the HSKP hierarchy with the theory of Darboux transformations. The fact that supersymmetry has an intriguing relation with the theory of Darboux transformations is a well established one. For instance, the classical problem of factorizing the Schrödinger operator $`_x^2+u(x)`$ into first order factors gives rise to a super algebra (see , §2, and references quoted therein). Another signal of this fact comes from Mulase’s paper , where it was shown how the modified KP equation can be obtained from the (Manin–Radul) SKP by means of a process of elimination of odd variables. We shall show that HSKP provides a natural framework to discuss such issues. We shall use the geometrical setting of , where the classical subject of Darboux–Bäcklund transformations and Miura–like maps is approached in a rather unconventional way that can be summarized as follows. Instead of searching directly for a symmetry of an evolutionary equation $`X`$ defined on a manifold $`M`$, one tries to find a covering, that is another evolutionary equation, defined on a bigger manifold $`N`$, related to $`X`$ by two maps $`\pi ,\sigma :NM`$, such that $$X=\pi _{}Y=\sigma _{}Y.$$ (1.8) In a covering for the KP hierarchy, called Darboux–KP (DKP) hierarchy, was constructed as a hierarchy defined on the phase space of pairs of Laurent series $`(h,a)`$ with suitable asymptotics and the (generalized) Miura transformation was defined to be $`h\sigma (a,h)=h+a_x/a.`$ Another application of this formalism has been used in to linearize the equations (the so called Central System (CS)) induced by the flows (1.4) on the currents $`H^{(j)}`$. This method has been exploited (see ) to construct a wide class of solutions of KP admitting a polynomial $`\tau `$–function. In this paper we will consider the supersymmetric counterpart of the geometric theory of the Darboux transformations. Firstly, we will show that the DKP system is actually the bosonic content of the even flows of HSKP (and hence of JSKP). Then we will construct a Darboux covering for HSKP and related “rational” reductions. Finally, we will use the technique of Darboux covering to linearize the Super analogue SCS of CS, exploiting this result in a quite explicit description of a wide class of non trivial solutions of HSKP of rational type. The detailed plan of the paper is the following: in Section 2, after having briefly recalled the basis of the (bi)hamiltonian set up for the KP theory, we will introduce the phase space for HSKP and define the hierarchy. In Section 2.2 we will discuss the fundamental properties of HSKP, and we will introduce the Super Central System SCS as the dynamical system obeyed by the currents of the theory when the Faà di Bruno generator $`\widehat{h}`$ evolves along HSKP. In Section 2.3 we will show how solutions of HSKP can be obtained from solutions of SCS, and in Section 2.4 we will briefly discuss how HSKP can be seen as a particular form of the Jacobian SKP hierarchy of Mulase and Rabin, by comparing the wave functions associated with the two theories. In Section 2.5, we will show how a super extension of the KdV equation can be obtained as a suitable reduction of HSKP; this will give us a concrete clue to the rest of the paper. Indeed, from Section 3 onwards we will turn our attention to the method of Darboux coverings. We will first recall the setting of the ordinary bosonic case, and then identify the bosonic part of HSKP with the DKP hierarchy of . We will also point out the specific form of the generalized Miura transformation. Furthermore, we will construct a Darboux covering of HSKP, and briefly discuss some reductions of the latter. Finally, in Section 4 we will show how the equations can be explicitly linearized, and discuss a specific class of solutions depending rationally on a finite number of times. ## 2 The GZ approach to KP and its supersymmetric extension The technique that plays a prominent role in the bi-Hamiltonian approach to KP is the Gelfan’d Zakharevich method of Poisson pencils to construct integrable Hamiltonian systems . In such a scheme one considers a manifold $`M`$ endowed with a pencil $`P_\lambda =P_1\lambda P_0`$ of Poisson structures, and studies the Casimir functions of the pencil. Such a Casimir function $`H_\lambda `$ is a (non-constant) function on $`M`$, which depends also on the parameter $`\lambda `$, such that $`P_\lambda dH_\lambda =0`$ for every value of $`\lambda `$. When $`M`$ is an $`2n+1`$–dimensional manifold endowed with a Poisson pencil of maximal rank, $`P_\lambda `$ has a unique Casimir function $`H_\lambda `$, which is a polynomial in $`\lambda `$ of degree $`n`$, $$H_\lambda =H_0\lambda ^n+H_1\lambda ^{n1}+\mathrm{}+H_n.$$ Its leading coefficient $`H_0`$ is the Casimir of $`P_0`$, while the “constant term” $`H_n`$ is the Casimir of $`P_1`$. The coefficients $`H_j`$ satisfy the recurrence relations $$P_1dH_{j+1}=P_0dH_j$$ and therefore are in involution with respect to all the brackets of the pencil. In the realm of infinite dimensional systems, the KdV theory is perhaps the best known prototype of a GZ hierarchy . Here, the manifold $`M`$ is the space of $`C^{\mathrm{}}`$ functions on the circle $`S^1`$, and the Poisson pencil, given as a one-parameter family of skew-symmetric maps from the cotangent to the tangent bundle, reads $$\dot{u}=(P_\lambda )_uv=\frac{1}{2}v_{xxx}+2(u+\lambda )v_x+u_xv,$$ where $`x`$ is a coordinate on $`S^1`$, $`u`$ represents a point of $`M`$, and $`\dot{u}`$ and $`v`$ are respectively a vector and a a covector at $`u.`$ It turns out that if $`h`$ and $`v`$ are series in $`z=\sqrt{\lambda }`$ of the form $$h(z)=z+\underset{j>0}{}h_jz^j,v=1+\underset{l>0}{}v_l/z^{2l}$$ (2.1) that provide the unique solutions of the Riccati system $$\{\begin{array}{cc}& h_x+h^2=u+z^2\hfill \\ & 1/2v_x+hv=z\hfill \end{array},$$ then $`v(z)`$ is the series representing the differential of the Casimir function of the Poisson pencil of KdV, which, in turn, is given by the integral $$H_\lambda =2z_{S^1}h𝑑x.$$ (2.2) The GZ hierarchy associated with $`H_\lambda `$ on $`M`$ admits several representations. The one we are interested in can be expressed by saying that the local Hamiltonian density $`h(z)`$ must obey local conservation laws of the form $$\frac{}{t_j}h=_xH^{(j)},$$ where the “current densities” $`H^{(j)}`$ are given by $$H^{(2j)}=\lambda ^j\text{ and }H^{(2j+1)}=\frac{1}{2}(\lambda ^jv)_{+,x}+h(\lambda ^jv)_+,$$ (2.3) the subscript $`+`$ meaning to take the positive part of the expansion in powers of $`z`$. Equation (2.3) can also be written as $$H^{(2j+1)}=z^{2j}\left(\frac{1}{2}v_x+hv\right)+\frac{1}{2}(z^{2j}v)_{,x}h(z^{2j}v)_{}$$ or $$H^{(2j+1)}=\underset{l=1}{\overset{j}{}}\left[\frac{1}{2}v_{jl,x}(z^{2l}1)+v_{jl}(z^{2l}h)\right].$$ The first of these two expressions shows that $`H^{(j)}=z^j+O(z^1)`$, since by the second Riccati equation above we have $`z^{2j}\left(\frac{1}{2}v_x+hv\right)=z^{2j+1}.`$ The interpretation of the second entails that the currents $`H^{(j)}`$ are the unique combinations $$H^{(j)}=\underset{k=0}{\overset{j}{}}c_k^jh^{(k)}$$ of the Faà di Bruno iterates $`h^{(j)}`$ of $`h^{(0)}=1`$ at $`h`$, defined by $$h^{(j+1)}=(_x+h)h^{(j)},$$ (2.4) that admit the asymptotic expansion $`H^{(j)}=z^j+O(z^1)`$. The relevance of this result is that the currents $`H^{(j)}`$ can be constructed without requiring that $`h`$ is a solution of the Riccati equation. Therefore one can define the KP hierarchy as follows. ###### Definition 2.1 Let $`h`$ be a monic (formal) Laurent series in $`z^1`$ $$h:=z+\underset{j>0}{}h_jz^j,$$ whose coefficients $`h_j`$ belong to $`C^{\mathrm{}}(S^1)`$, and consider its Faà di Bruno iterates $`h^{(j)}`$. Denote by $`W`$ the span over $`C^{\mathrm{}}(S^1)`$ of the order Faà di Bruno iterates (or monomials) $$W:=\mathrm{span}_{C^{\mathrm{}}(S^1)}\{h^{(k)},k0\}$$ and introduce the “current densities” $`H^{(k)}`$ by requiring them to be the unique elements of $`W`$ of the form $$H^{(k)}=z^k+\underset{j>0}{}H_j^kz^j.$$ The KP hierarchy is defined to be the set of conservation laws $$\frac{}{t_k}h=_xH^{(k)}.$$ (2.5) We observe that $`H^{(1)}=h`$, so we can identify the first time $`t_1`$ with $`x`$. Moreover, it can be proven that this definition is completely equivalent to the one given in the framework of pseudo-differential operators (see, e. g., and references quoted therein). We end this review of the bihamiltonian set-up of the KP hierarchy with the notion of the Central System (CS). The operators $`_{t_k}+H^{(k)}`$ satisfy the invariance condition $$\left(_{t_k}+H^{(k)}\right)WW$$ and the commutativity property $`[_{t_k}+H^{(k)},_{t_j}+H^{(j)}]=0`$. This entails that along the KP flows (2.5) the currents $`H^{(k)}`$ satisfy the following evolutionary equation: $$\frac{}{t_j}H^{(k)}+H^{(j)}H^{(k)}=H^{(j+k)}+\underset{l=1}{\overset{k}{}}H_l^jH^{(kl)}+\underset{l=1}{\overset{j}{}}H_l^kH^{(jl)}.$$ (2.6) ###### Definition 2.2 Let $``$ be the space of sequences $`\{H^{(0)},H^{(1)},H^{(2)},H^{(3)},\mathrm{}\}`$ where $`H^{(0)}=1`$ and the $`H^{(j)}`$’s are of the form: $$H^{(j)}=z^j+\underset{l1}{}\frac{H_l^j}{z^l},j1$$ The Central System is the hierarchy of dynamical system on $``$ defined by equation (2.6). It turns out that the vector fields of CS commute among themselves . In the next Section we define the extension to the ($`N=1`$) supersymmetric case of the constructions herewith outlined. This will lead us to the definition of a supersymmetric extension of the KP hierarchy. We will refer to such a SKP theory as Hamiltonian Super KP, to keep track of its (albeit remote) hamiltonian origins. Later (see Section 2.4) we will show how to identify our HSKP with the Jacobian super KP hierarchy of Mulase and Rabin ) . ### 2.1 The definition of the Hamiltonian super KP hierarchy Let us start by fixing some notations (see, e.g., for more details on supergeometry), to be used throughout the paper. We denote by $`\mathrm{\Lambda }`$ a generic Grassmann algebra over $``$. This is required by functorial properties of supersymmetry , but in this work it will play a spectator role, and can be thought of as a fixed algebra. We supplement the bosonic spectral parameter $`z`$ with its fermionic “super-partner” $`\theta `$, and replace the circle by its super analog $`S^{1|1}`$ endowed with coordinates $`x`$, $`\phi `$. To simplify notations we call $`B_{[x\phi ]}`$ the ring $`C^{\mathrm{}}(S^{1|1},\mathrm{\Lambda })`$ of smooth functions on $`S^{1|1}`$ with values in $`\mathrm{\Lambda }`$ (or a suitable “version” of it, like the space of $`\mathrm{\Lambda }`$-valued functions on $`^{1|1}`$ vanishing at infinity, or even the space $`[[x,\phi ]]\mathrm{\Lambda }`$ of formal series in $`x,\phi `$.) Finally, we denote by $`\overline{f}`$ the parity of a homogeneous element $`f`$, e.g. $`\overline{z}=\overline{x}=0`$, $`\overline{\theta }=\overline{\phi }=1`$. The Hamiltonian super KP hierarchy is defined in terms of the super Faà di Bruno generator $`\widehat{h}`$ and the odd derivation operator $`\delta :=_\phi +\phi _x`$ taking the place of $`_x.`$ The generator $`\widehat{h}`$ is an odd formal Laurent series in $`z^1`$ and $`\theta `$ with coefficients in $`B_{[x\phi ]}`$ of the form $$\widehat{h}(z,\theta ;x,\phi ):=\nu (z;x)+\theta a(z;x)+\phi h(z;x)+(\theta \phi )\psi (z;x),$$ (2.7) We specify exactly the content of the components $`\nu `$, $`a`$, $`h`$ and $`\psi `$, by analogy with the KP formalism, requiring that: 1. All equations are homogeneous with respect to the grading specified by $$[\theta ]=\frac{1}{2},[z]=1,[\phi ]=\frac{1}{2},[x]=1,[\widehat{h}]=[\delta ]=\frac{1}{2},[t_j]=\frac{j}{2},$$ and no field of negative weight enters the theory. 2. It is possible to identify the second time $`t_2`$ with $`x`$.<sup>1</sup><sup>1</sup>1 The relation of $`t_1`$ with $`\phi `$ will be however much subtler. We will discuss it in Section 2.3. 3. There exist suitable “super current densities” $`\widehat{H}^{(k)}`$, with asymptotic behavior $$\widehat{H}^{(2j+p)}z^j\theta ^p+O(1/z),j,p\{0,1\}.$$ It turns out that these requirements can be satisfied if the following simple and convenient choices are made: $`a`$ is holomorphic in $`z^1`$ with constant and invertible zeroth order coefficient (which we assume to be equal to $`1`$): $`a(z;x):=1+_{j>0}a_j(x)z^j.`$ $`h`$ has the usual form $`h(z;x):=z+_{j>0}h_j(x)z^j.`$ $`\nu `$ and $`\psi `$ are of the form $$\nu (z;x):=\underset{j>0}{}\nu _j(x)z^j,\psi (z;x):=\underset{j>0}{}\psi _j(x)z^j.$$ With the super Faà di Bruno generator $`\widehat{h}`$ we associate (for $`k`$) its iterates $$\{\begin{array}{c}\widehat{h}^{(k+1)}:=(\delta +\widehat{h})\widehat{h}^{(k)}\hfill \\ \widehat{h}^{(0)}:=1\hfill \end{array}.$$ The following lemma shows that the even Faà di Bruno iterates, apart from their nilpotent components, are essentially the usual Faà di Bruno monomials. ###### Lemma 2.1 Let $`\widehat{f}:=\delta (\widehat{h})=h\theta \psi +\phi \nu _x(\theta \phi )a_x`$. Then, for any $`k`$ $$\{\begin{array}{c}\widehat{h}^{(2k+2)}=(_x+\widehat{f})\widehat{h}^{(2k)}=(_x+\widehat{f})^{k+1}1\hfill \\ \widehat{h}^{(2k+3)}=(_x+\widehat{f})\widehat{h}^{(2k+1)}=(_x+\widehat{f})^{k+1}\widehat{h}\hfill \end{array}.$$ (2.8) Proof. We have $$(\delta +\widehat{h})^2=\delta ^2+\delta (\widehat{h})\widehat{h}\delta +\widehat{h}\delta +\widehat{h}^2=\delta ^2+\delta (\widehat{h})=_x+\widehat{f},$$ so we get $$\widehat{h}^{(2k+2)}=(\delta +\widehat{h})^2\widehat{h}^{(2k)}=(_x+\widehat{f})\widehat{h}^{(2k)}=(\delta +\widehat{h})^{2k+2}1=(_x+\widehat{f})^{k+1}1$$ and $$\widehat{h}^{(2k+3)}=(\delta +\widehat{h})^2\widehat{h}^{(2k+1)}=(_x+\widehat{f})\widehat{h}^{(2k+1)}=(\delta +\widehat{h})^{2k+2}\widehat{h}=(_x+\widehat{f})^{k+1}\widehat{h}.$$ $`\mathrm{}`$ For later use we express the Faà di Bruno iterates as $$\{\begin{array}{c}\widehat{h}^{(2k)}=h^{(k)}\theta \psi ^{(k)}+\phi \omega ^{(k)}(\theta \phi )b^{(k)}\hfill \\ \widehat{h}^{(2k1)}=\nu ^{(k)}+\theta a^{(k)}+\phi d^{(k)}+(\theta \phi )\chi ^{(k)}\hfill \end{array},$$ (2.9) where the components are Laurent series of the form $$\begin{array}{cc}\nu ^{(k)}=_{j>0}\nu _j^{(k)}z^{kj1},\hfill & h^{(k)}=z^k+_{j>0}h_j^{(k)}z^{kj1}\hfill \\ a^{(k)}=z^{k1}+_{j>0}a_j^{(k)}z^{kj1},\hfill & \psi ^{(k)}=_{j>0}\psi _j^{(k)}z^{kj1}\hfill \\ d^{(k)}=z^k+_{j>0}d_j^{(k)}z^{kj1},\hfill & \omega ^{(k)}=_{j>0}\omega _j^{(k)}z^{kj1}\hfill \\ \chi ^{(k)}=_{j>0}\chi _j^{(k)}z^{kj1},\hfill & b^{(k)}=_{j>0}b_j^{(k)}z^{kj1}\hfill \end{array}$$ and can be computed by recurrence according to the rules displayed in Table 1 By analogy with the KP hierarchy, we introduce the space $$W_{B_{[x\phi ]}}:=\mathrm{span}_{B_{[x\phi ]}}\{\widehat{h}^{(k)},k\}$$ (2.10) and prove the existence of the super currents with the desired asymptotic behavior. ###### Proposition 2.2 Let $`\widehat{h}`$ and $`W_{B_{[x\phi ]}}`$ be defined as in equations (2.7) and (2.10). There exists a basis $`\{\widehat{H}^{(k)},k\}`$ of $`W_{B_{[x\phi ]}}`$ with $$\{\begin{array}{c}\widehat{H}^{(2k)}=z^k+_{j>0}\left(\widehat{H}_{0,j}^{2k}(x,\phi )z^j+\widehat{H}_{1,j}^{2k}(x,\phi )\theta z^j\right)\hfill \\ \widehat{H}^{(2k+1)}=\theta z^k+_{j>0}\left(\widehat{H}_{0,j}^{2k+1}(x,\phi )z^j+\widehat{H}_{1,j}^{2k+1}(x,\phi )\theta z^j\right)\hfill \end{array}.$$ Proof. By definition of $`W_{B_{[x\phi ]}}`$ we see that $$\{\begin{array}{c}\widehat{H}^{(0)}=1\hfill \\ \widehat{H}^{(1)}=\widehat{h}^{(1)}\phi \widehat{h}^{(2)}\hfill \\ \widehat{H}^{(2)}=\widehat{h}^{(2)}\hfill \end{array}.$$ (2.11) The others can be computed recursively: suppose we have defined $`\widehat{H}^{(j)}`$ for $`0j<k`$; if $`k=2n`$ is even then $$\widehat{H}^{(k)}=\widehat{h}^{(k)}\underset{j=1}{\overset{n1}{}}\left(h_j^{(n)}+\phi \omega _j^{(n)}\right)\widehat{H}^{(k2j2)}\underset{j=1}{\overset{n1}{}}\left(\psi _j^{(n)}+\phi b_j^{(n)}\right)\widehat{H}^{(k2j1)},$$ (2.12) while if $`k=2n1`$ is odd then $$\begin{array}{c}\widehat{H}^{(k)}=\widehat{h}^{(k)}\phi \widehat{h}^{(k+1)}_{j=1}^{n1}\left(\nu _j^{(n)}+\phi (d_j^{(n)}h_j^{(n)})\right)\widehat{H}^{(k2j1)}\hfill \\ _{j=1}^{n1}\left(a_j^{(n)}+\phi (\chi _j^{(n)}\psi _j^{(n)})\right)\widehat{H}^{(k2j)}\hfill \end{array}.$$ (2.13) We have thus prepared all the “ingredients” needed for the following ###### Definition 2.3 (HSKP) Let $`\widehat{h}`$ be defined by (2.7), and compute its Faà di Bruno iterates $`\widehat{h}^{(k)}`$ and the basis $`\{\widehat{H}^{(k)},k0\}`$ of $`W_{B_{[x\phi ]}}`$ as explained in Proposition 2.2. The Hamiltonian super KP hierarchy is the set of “super conservation laws” $$_{t_k}\widehat{h}=(1)^k\delta \widehat{H}^{(k)},k>0.$$ (2.14) Notice that, according to the last line of (2.11), one has $`\widehat{H}^{(2)}=\widehat{h}^{(2)}=\delta \widehat{h}`$. Hence the $`t_2`$ equation of motion of HSKP is $$_{t_2}\widehat{h}=\delta \widehat{H}^{(2)}=\delta (\delta \widehat{h})=_x\widehat{h},$$ that is, indeed, $`t_2`$ can be identified with $`x`$. ### 2.2 The super central system The first property to be verified is the compatibility of the evolution equations (2.14). While checking this, we shall find that the hierarchy has some very useful properties which allow us to describe it as producing flows on the super universal Grassmannian. This will be accomplished by the introduction of a dynamical system tightly connected with HSKP. Let us first work out some simple consequences of the definition of the hierarchy. The evolution equations (2.14) are simply the super-commutativity conditions $$[\delta +\widehat{h},_{t_k}+\widehat{H}^{(k)}]=0$$ and imply that $$\left(_{t_k}+\widehat{H}^{(k)}\right)W_{B_{[x\phi ]}}W_{B_{[x\phi ]}}.$$ (2.15) Indeed, $$\begin{array}{cc}\hfill \left(_{t_k}+\widehat{H}^{(k)}\right)& \widehat{h}^{(l)}=(_{t_k}+\widehat{H}^{(k)})(\delta +\widehat{h})^l1\hfill \\ & =(1)^{kl}\left(\delta +\widehat{h}\right)^l\left(_{t_k}+\widehat{H}^{(k)}\right)1=(1)^{kl}\left(\delta +\widehat{h}\right)^l\widehat{H}^{(k)}\hfill \end{array}$$ and by definition $`\widehat{H}^{(k)}W_{B_{[x\phi ]}}`$, $`(\delta +\widehat{h})W_{B_{[x\phi ]}}W_{B_{[x\phi ]}}`$. In turn, this implies the “abelian zero curvature” equation: $$_{t_j}\widehat{H}^{(k)}=(1)^{jk}_{t_k}\widehat{H}^{(j)},$$ (2.16) as the following simple argument shows. Denote by $`V_{B_{[x\phi ]}}`$ the space of Laurent series in $`z^1`$ and $`\theta `$ with coefficients in $`B_{[x\phi ]}`$ and by $`V_{B_{[x\phi ]}}^{}`$ its subspace of formal power series without “constant term”, i.e. starting from $`z^1`$ and $`\theta z^1`$. We have the decomposition $$V_{B_{[x\phi ]}}=W_{B_{[x\phi ]}}V_{B_{[x\phi ]}}^{}.$$ (2.17) Then, by Equation (2.15), $`_{t_j}\widehat{H}^{(k)}`$ is the $`V_{B_{[x\phi ]}}^{}`$-component of $`\widehat{H}^{(j)}\widehat{H}^{(k)}`$, while $`_{t_k}\widehat{H}^{(j)}`$ is the $`V_{B_{[x\phi ]}}^{}`$-component of $`\widehat{H}^{(k)}\widehat{H}^{(j)}=(1)^{jk}\widehat{H}^{(j)}\widehat{H}^{(k)}`$, thus proving equations (2.16). Thanks to this property and the commutativity of the operators $`\delta `$ and $`_{t_j}`$, we obtain the compatibility of the evolution equations $$_{t_j}_{t_k}\widehat{h}=(1)^{j+k}\delta _{t_j}\widehat{H}^{(k)}=(1)^{jk+j+k}\delta _{t_k}\widehat{H}^{(j)}=(1)^{jk}_{t_k}_{t_j}\widehat{h}.$$ This result finally entails the supercommutativity of the operators $`_{t_k}+\widehat{H}^{(k)}`$, $$[_{t_j}+\widehat{H}^{(j)},_{t_k}+\widehat{H}^{(k)}]=0.$$ We notice that it is possible to describe the theory in terms of the super-currents $`\widehat{H}^{(k)}`$’s only, avoiding the introduction of the super-space variables $`x`$ and $`\phi `$ and the super-derivative $`\delta `$ which up to now played a special role. It is by doing this that the super universal Grassmannian arises. Let us first of all recall its definition . Denote by $`V:=\mathrm{\Lambda }((z^1))\mathrm{\Lambda }((z^1))\theta `$ the quotient ring of the ring of formal power series in $`z^1`$ and $`\theta `$ over<sup>2</sup><sup>2</sup>2To make contact with the previous definitions, notice that $`V_{B_{[x\phi ]}}=V_\mathrm{\Lambda }B_{[x\phi ]}`$. $`\mathrm{\Lambda }`$, and let $`V_+:=\mathrm{\Lambda }[z,\theta ]`$, $`V_{}:=\mathrm{\Lambda }[[z^1,\theta ]]z^1.`$ $`V`$ has a natural filtration $$\mathrm{}V_{j1}V_jV_{j+1}\mathrm{},$$ where $`V_j=z^{j+1}V_{}`$, which makes it and its $`\mathrm{\Lambda }`$-submodule $`V_+`$ complete topological spaces. Then, the super Grassmannian $`SGr_\mathrm{\Lambda }:=SGr_\mathrm{\Lambda }(V,V_+)`$ is the set of closed free $`\mathrm{\Lambda }`$-submodules $`W`$ of $`V`$ which are compatible with $`V_+`$ in the sense that the restriction $`\pi _W`$ of the projection $`\pi _W:VV_+`$ to $`W`$ is a Fredholm operator, i.e. its kernel (respectively cokernel) is a $`\mathrm{\Lambda }`$-submodule (respectively a $`\mathrm{\Lambda }`$-quotient module) of a finite rank free $`\mathrm{\Lambda }`$-module. As usual , $`SGr_\mathrm{\Lambda }`$ is the disjoint union of the denumerable set of its components labelled by the index $`i_W`$ of $`\pi _W`$. We exploit our formulæ and the concept of super universal Grassmannian by means of the following ###### Definition 2.4 (SCS) Let $`M`$ be the set of sequences $`\{\widehat{H}^{(k)}\}_{k0}`$ of formal Laurent series with coefficients in $`\mathrm{\Lambda }`$ admitting the following expansion in $`z`$: $$\{\begin{array}{c}\widehat{H}^{(2k)}=z^k+_{j>0}\left(\widehat{H}_{0,j}^{2k}z^j+\widehat{H}_{1,j}^{2k}\theta z^j\right)\hfill \\ \widehat{H}^{(2k+1)}=\theta z^k+_{j>0}\left(\widehat{H}_{0,j}^{2k+1}z^j+\widehat{H}_{1,j}^{2k+1}\theta z^j\right)\hfill \end{array},$$ with $`\widehat{H}^{(0)}=1`$, $`\overline{\widehat{H}^{(k)}}=kmod2`$, and let $$W=\mathrm{span}_\mathrm{\Lambda }\{\widehat{H}^{(k)},k0\}V.$$ It is not difficult to realize that $`M`$ is isomorphic to the big cell of the $`0|0`$ component of $`SGr_\mathrm{\Lambda }`$ (i.e. the open subset where $`\mathrm{ker}\pi _W=\mathrm{coker}\pi _W=0`$), an explicit isomorphism being given by the map $$\{\widehat{H}^{(k)}\}_{k0}\mathrm{span}_\mathrm{\Lambda }\{\widehat{H}^{(k)},k0\}.$$ The super central system is the dynamical system defined on $`M`$ by requiring that $$(_{t_j}+\widehat{H}^{(j)})WW$$ or, equivalently, $$_{t_j}\widehat{H}^{(k)}=\pi _{}(\widehat{H}^{(j)}\widehat{H}^{(k)}),$$ (2.18) where $`\pi _{}:VV_{}`$ is the projection of $`V=V_{}V_+`$ onto $`V_{}`$ parallel to $`V_+`$. By comparing coefficients in the formulation (2.18) of the super central system, we can explicitly write its evolution equations (see Table 2). Notice that the SCS is to be thought of as a system of $`_2`$-graded ordinary differential equations. ### 2.3 HSKP as a “reduction” of SCS The Hamiltonian super KP hierarchy (2.14) can be obtained from SCS by “spatialisation”. This procedure will be used in Section 4 to produce solutions of HSKP starting from solutions of SCS. A spatialization of a hierarchy of dynamical systems $`X`$ is a process, (see, e.g., ), consisting in the projection of $`X`$ onto the space $`\widehat{𝒬}_j`$ of solutions of its $`j`$-th flow. More informally, it boils down to interpret a distinguished flow parameter as a space coordinate, and allows to interpret the dynamical system as a system of PDEs. In the ordinary KP case, spatialization with respect to the time $`t_1=x`$ simply amounts to identify $`t_1`$ with $`x`$ (or better, substitute $`t_1=x+t_1`$ in the solutions of CS). This procedure yields that $`h(x)=H_{|_{t_1=t_1+x}}^{(1)}`$ is a solution to the KP equations. In the super case, we want to consider the projection of SCS to the space $`\widehat{𝒬}_2`$ of solutions of its second flow i.e. the space of orbits of $`_2`$. Essentially, we have to consider $`k=1`$ and interpret the first two families of equations of motion reported in Table 2 as recursive definitions of the currents, as differential polynomials (in the space variable $`x=t_2`$) of the generators $`\widehat{H}^{(1)}`$ and $`\widehat{H}^{(2)}`$. With respect to the bosonic case, there is a subtlety, connected with the relation of the first time $`t_1`$ of SCS with the fermionic partner $`\phi `$ of $`x`$. Observe that, by the definition of the super central system, we have $`_{t_j}\widehat{H}^{(k)}=(1)^{jk}_{t_k}\widehat{H}^{(j)}`$. Notice in particular that $`_1\widehat{H}^{(1)}=0`$. Now, for $`k>1`$ $$(1)^k(_1+t_1_2)\widehat{H}^{(k)}=_{t_k}(\widehat{H}^{(1)}+t_1\widehat{H}^{(2)}),$$ (2.19) suggesting that in order to get HSKP (and solutions thereof) we should put $`\widehat{h}=\widehat{H}^{(1)}+t_1\widehat{H}^{(2)}`$, $`t_2=x`$ and $`t_1=\phi `$. This is “almost true”, but one must pay attention to the order in which these identifications are performed. Indeed, plugging $`k=1`$ in the left hand side of (2.19), we get $$(_1+t_1_2)\widehat{H}^{(1)}=t_1_1\widehat{H}^{(2)}_1(\widehat{H}^{(1)}+t_1\widehat{H}^{(2)}),$$ which is unconsistent. The right way to proceed is the following. Starting from a solution of SCS, which depends on the times $`(t_1,t_2,\mathrm{})=𝐭`$, one first replaces in the currents $`\widehat{H}^{(j)}(𝐭)`$ the times $`t_1`$ with $`t_1+\phi `$ and $`t_2`$ with $`x`$, then one defines $$\widehat{h}(x,\phi ;𝐭):=\widehat{H}^{(1)}(t_1+\phi ,x,\mathrm{})+\phi \widehat{H}^{(2)}(t_1+\phi ,x,\mathrm{}).$$ (2.20) Since $`_1\widehat{H}^{(k)}=_\phi \widehat{H}^{(k)}`$ for any $`k`$, and taking also into account (2.19), now we have that the field $`\widehat{h}(x,\phi ;𝐭)`$ is a solution of $`_{t_k}\widehat{h}(x,\phi ;𝐭)=(1)^k\delta \widehat{H}^{(k)}`$, i.e. that it indeed satisfies the HSKP hierarchy. Observe that $`\widehat{H}^{(2)}=\widehat{h}^{(2)}=(\delta \widehat{h})`$; in fact $`\widehat{h}^{(2)}`$ $`=`$ $`\delta (\widehat{H}^{(1)}+\phi \widehat{H}^{(2)})=_\phi \widehat{H}^{(1)}+\phi _x\widehat{H}^{(1)}+_\phi (\phi \widehat{H}^{(2)})`$ $`=`$ $`\phi _\phi \widehat{H}^{(2)}+(\widehat{H}^{(2)}\phi _\phi \widehat{H}^{(2)})=\widehat{H}^{(2)}.`$ Finally one has that $`(_{t_j}+\widehat{H}^{(j)})WW`$ implies $`(\delta +\widehat{h})W_{B_{[x\phi ]}}W_{B_{[x\phi ]}}`$. ### 2.4 The connection with the JSKP of Mulase and Rabin In this section we will show that HSKP is equivalent to the Mulase–Rabin Jacobian Super KP hierarchy. Although it seems conceivable from the supercommutativity of the flows, such an identification is somewhat subtle. Our essential tool will be the introduction of two wave or Baker–Akhiezer functions associated with HSKP. The zero curvature condition (2.16) implies the existence of a first wave function $`\mathrm{\Phi }`$ satisfying $$\frac{}{t_k}\mathrm{\Phi }=\widehat{H}^{(k)}\mathrm{\Phi }.$$ (2.21) Now we perform the following “trick”, whose meaning will be discussed in Remark 2.1. We define “enhanced” currents $`\widehat{}^{(j)}`$ by the formula: $$\widehat{}^{(j)}=\widehat{H}^{(j)}+(1)^{j+1}\phi \widehat{H}^{(j)}𝑑z𝑑\theta .$$ (2.22) In words, the difference between $`\widehat{}^{(j)}`$ and $`\widehat{H}^{(j)}`$ is (up to a sign) the $`\theta `$ component of the residue in $`z`$ of $`\widehat{H}^{(j)}`$, multiplied by $`\phi `$; in the sequel we will denote it as $$\widehat{}^{(j)}\widehat{H}^{(j)}=(1)^{j+1}\phi C_j.$$ The zero curvature condition on the currents $`\widehat{H}^{(j)}`$ implies that the enhanced currents satisfy the analogue condition $$\frac{\widehat{}^{(j)}}{t_n}=(1)^{jn}\frac{\widehat{}^{(n)}}{t_j},$$ (2.23) and so guarantees the existence of an enhanced wave function $`\mathrm{\Psi }`$ satisfying $$\frac{}{t_n}\mathrm{\Psi }=\widehat{}^{(n)}\mathrm{\Psi }.$$ (2.24) The wave function $`\mathrm{\Psi }`$ is readily seen to be related to the $`\mathrm{\Phi }`$–wave function by the formula: $$\mathrm{\Psi }=\mathrm{\Phi }\mathrm{exp}\left(\phi _{𝐭_0^{ev}}^{𝐭^{ev}}\underset{n}{}C_{2n}ds_{2n}\right).$$ (2.25) We now consider the logarithmic derivative $$\widehat{𝔥}=\delta \mathrm{\Psi }/\mathrm{\Psi }.$$ (2.26) It is related with the Faà di Bruno generator $`\widehat{h}`$ by $$\widehat{𝔥}=\widehat{h}_{𝐭_0^{ev}}^{𝐭^{ev}}\underset{n}{}C_{2n}ds_{2n}.$$ (2.27) Actually, since $`C_2`$ is readily seen to be $`\psi _1`$, we can say that $`\widehat{𝔥}`$ is a $`^{}[z,\theta ]`$–valued superfield of the form $$\widehat{𝔥}=\theta a+\stackrel{~}{\nu }+\phi h+(\theta \phi )\psi ,$$ (2.28) where now $$\stackrel{~}{\nu }=\nu _0+\underset{j1}{}\frac{\nu _j}{z^j},\text{ with }\nu _{0x}=\psi _1.$$ (2.29) We notice that, by a straightforward supersymmetric extension of standard properties of the Faà di Bruno procedure, since $`\widehat{𝔥}`$ differs from $`\widehat{h}`$ by a zero order term in $`z`$, one can write the Faà di Bruno iterates $`\widehat{𝔥}^{(j)}`$ of $`\widehat{𝔥}`$ as a linear (over $`B_{[x\phi ]}`$) combination of the iterates $`\widehat{h}^{(j)}`$ we have been using so far (and conversely). To grasp this fact, one simply has to notice that Lemma 2.1 holds irrespectively of the fact that $`\nu _0`$ vanishes. Since $`\widehat{f}^{}=\delta \widehat{𝔥}=\widehat{f}+\delta \nu _0`$, we have (using induction) that $$\begin{array}{cc}\hfill (_x+\widehat{f}^{})& \alpha _i\widehat{h}^{(i)}=\left(_x+\widehat{f}+\delta \nu _0\right)(\alpha _i\widehat{h}^{(i)})\hfill \\ & =\alpha _i\widehat{h}^{(i+2)}+\alpha _{ix}\widehat{h}^{(i)}+(\delta \nu _0\alpha _i)\widehat{h}^{(i)}.\hfill \end{array}$$ Summing up, we see that we can express the enhanced currents as linear combinations of the Faà di Bruno iterates of $`\widehat{𝔥}`$: $$\widehat{}^{(j)}=\underset{m}{}\mathrm{\Gamma }_m^j\widehat{𝔥}^{(m)}.$$ Thanks to the obvious equality $`\widehat{𝔥}^{(j)}\mathrm{\Psi }=\delta ^j\mathrm{\Psi }`$, we can write equation (2.24) as follows: $$_{t_j}\mathrm{\Psi }=\widehat{}^{(j)}\mathrm{\Psi }=\underset{l}{}\mathrm{\Gamma }_l^j\widehat{𝔥}^{(l)}\mathrm{\Psi }=B_j\mathrm{\Psi },$$ (2.30) where $`B_j`$ is a super differential operator of order $`j`$ and of parity $`jmod2`$. This equation is the bridge between HSKP and JSKP. The latter is usually formulated within the theory of super pseudo-differential operators. A discussion of such a topic is outside the size of this paper (see, e.g., for details). We need the following lemma, whose proof is a straightforward computation: ###### Lemma 2.3 Let $`S`$ be a super pseudo-differential operator, with coefficients in $`B_{[x\phi ]}`$ of the form $$S=1+\underset{j>0}{}\left(u_j+\phi \xi _j\right)_x^j+\left(\eta _j+\phi w_j\right)\delta ^{(2j1)}.$$ (2.31) and let $$e(z,\theta ;x,\phi ,𝐭)=\mathrm{exp}\left(\phi \theta +zx+\underset{j>0}{}(t_{2j}z^j+t_{2j1}\theta z^{j1})\right)$$ be the vacuum wave function (in the terminology of ) for JSKP. Then, if $`\mathrm{\Psi }`$ is a Baker–Akhiezer function for JSKP obtained by dressing with $`S`$ the vacuum wave function $`e`$, $$\mathrm{\Psi }=Se,$$ (2.32) its logarithmic derivative $`\delta \mathrm{\Psi }/\mathrm{\Psi }`$ is a superfield of the form (2.28), satisfying the constraint $`\nu _{0x}+\psi _1=0`$. The JSKP equations can now be obtained by means of standard procedures in the theory of integrable systems. Indeed, taking the $`t_{2j}`$ and $`t_{2j1}`$ derivatives of the dressing relation (2.32), and taking also equation (2.30) into account, we have $$\begin{array}{cc}\hfill _{t_{2j}}\mathrm{\Psi }& =_{t_{2j}}(Se)=(_{t_{2j}}S)e+Sz^je=(_{t_{2j}}S)e+S\delta ^{2j}e\hfill \\ & =\left((_{t_{2j}}S)S^1+S\delta ^{2j}S^1\right)Se=\left((_{t_{2j}}S)S^1+S\delta ^{2j}S^1\right)\mathrm{\Psi }=B_{2j}\mathrm{\Psi }\hfill \end{array}$$ (2.33) and $$\begin{array}{cc}\hfill _{t_{2j1}}\mathrm{\Psi }& =_{t_{2j1}}(Se)=(_{t_{2j1}}S)e+S\theta z^{j1}e=(_{t_{2j1}}S)e+S(\delta ^{2j1}\phi \delta ^{2j})e\hfill \\ & =\left((_{t_{2j1}}S)S^1+S(\delta ^{2j1}\phi \delta ^{2j})S^1\right)\mathrm{\Psi }=B_{2j1}\mathrm{\Psi }.\hfill \end{array}$$ (2.34) Since $`(_jS)S^1=((_jS)S^1)_{}`$ is a purely pseudo-differential operator (i.e. it has no differential part) we get $$_{t_{2j}}S=(S\delta ^{2j}S^1)_{}S=(S_x^jS^1)_{}S$$ and $$_{t_{2j1}}S=(S(\delta ^{2j1}\phi \delta ^{2j})S^1)_{}S=(S_\phi _x^{j1}S^1)_{}S,$$ which are the equations that Mulase and Rabin defined for JSKP. ###### Remark 2.1 As we have anticipated, the introduction of the enhanced currents $`\widehat{}^{(j)}`$ is not a mere trick. To better understand their origin it is useful to reconsider our choices, namely the decomposition (2.17) of the space $`V_{B_{[x\phi ]}}`$ of formal Laurent series with coefficients in $`B_{[x\phi ]}`$ as the direct sum of the subspace generated by the Faà di Bruno monomials $`W_{B_{[x\phi ]}}`$ and the space $`V_{B_{[x\phi ]}}^{}`$ of formal power series without “constant term”, i.e. starting from $`z^1`$ and $`\theta z^1`$. If $`\pi _+:V_{B_{[x\phi ]}}W_{B_{[x\phi ]}}`$ is the projection associated with such a decomposition, then the currents $`\widehat{H}^{(k)}`$ are given by the formulas $$\widehat{H}^{(2j+p)}=\pi _+(z^j\theta ^p),j,p\{0,1\}.$$ Actually, associated with our geometrical datum, there is another natural choice. Indeed, one simply notice the fact that it is possible to extend the Faà di Bruno recursion relations (2.8) to negative values of the index $`j`$, and get a full Faà di Bruno basis $`\left\{\widehat{𝔥}^{(j)}\right\}_j`$ in $`V_{B_{[x\phi ]}}`$. The asymptotics of the Faà di Bruno basis is readily seen to be $$\widehat{𝔥}^{(2j)}z^j,\widehat{𝔥}^{(2j1)}\theta z^{j1}+\phi z^j.$$ (2.35) Hence we have another natural decomposition $$V_{B_{[x\phi ]}}=W_{B_{[x\phi ]}}W_{B_{[x\phi ]}}^{},$$ (2.36) where now $`W_{B_{[x\phi ]}}^{}`$ is the space generated by the Faà di Bruno iterates with negative index. If we call $$\pi _+^{}:V_{B_{[x\phi ]}}W_{B_{[x\phi ]}}$$ the projection associated with this new decomposition, then we have that the enhanced currents are given by $$\widehat{}^{(2j+p)}=\pi _+^{}(z^j\theta ^p),j,p\{0,1\}.$$ Actually, it is not surprising that the connection with the usual formulation of the theory by means of super pseudo-differential operators can be better described using the decomposition associated with the full Faà di Bruno basis, since $`\delta ^j\mathrm{\Psi }=\widehat{𝔥}^{(j)}\mathrm{\Psi }`$ and the projection $`\pi _+^{}`$ is adapted to the projection which kills the non-differential part of a super pseudodifferential operator. $`\mathrm{}`$ ###### Remark 2.2 It is actually possible to write an evolution equation of the form (2.30) for the first wave function $`\mathrm{\Phi }`$ as well. Indeed, since $`\widehat{h}\mathrm{\Phi }=\delta \mathrm{\Phi }`$ and $`\widehat{H}^{(j)}`$ is a linear combination of the Faà di Bruno iterates $`\widehat{h}^{(j)}`$ of $`\widehat{h}`$, we can read the equation $`_{t_j}\mathrm{\Phi }=\widehat{H}^{(j)}\mathrm{\Phi }`$ as $`_{t_j}\mathrm{\Phi }=B_j^{}\mathrm{\Phi }`$ where $`B_j^{}`$ is still a super differential operator of order $`j`$ and of parity $`jmod2`$. However, we can no more express $`\mathrm{\Phi }`$ as the result of the action of a generic dressing operator $`S`$ on the vacuum wave function $`e`$ of Lemma 2.3. Moreover, the HSKP equations are not compatible with $`\mathrm{\Phi }`$ having such an expression: even if $`\psi _1=0`$ at $`𝐭_0`$, this is no more true for the evolved field so $`0=\nu _{0x}\psi _1`$. $`\mathrm{}`$ ###### Remark 2.3 As we have seen, the connection between HSKP and JSKP is (albeit in a tricky way) a change of coordinates. The relation among the degrees of freedom $`(u_i,w_i,\xi _i,\eta _i)`$ of $`S`$ and the degrees of freedom $`(a_i,h_i,\nu _i,\psi _i)`$ (collected in $`(a(z),h(z),\nu (z)\psi (z))`$ as usual) of $`\widehat{𝔥}`$ is indeed the following: $$\begin{array}{c}\nu (z)\left(1+_{i>1}\frac{u_i}{z^i}\right)=\eta _1+\underset{i>1}{}\frac{\xi _i\eta _{i+1}}{z^i};\hfill \\ a(z)\left(1+_{i>1}\frac{u_i\eta _i\nu (z)}{z^i}\right)=1+\underset{i>1}{}\frac{u_i+w_i}{z^i};\hfill \\ h(z)\left(1+_{i>1}\frac{u_i}{z^i}\right)=z+u_1+\eta _1\nu (z)+\underset{i>1}{}\frac{u_{i,x}+u_{i+1}}{z^i}\hfill \\ +\nu (z)\left(_{i>1}\frac{\xi _i\eta _{i+1}}{z^i}\right);\hfill \\ \psi (z)\left(1+_{i>1}\frac{u_i}{z^i}\right)=a(z)\left(\eta _1+\underset{i>1}{}\frac{\eta _{i+1}\xi _i}{z^i}\right)+\nu (z)\underset{i>1}{}\frac{w_i}{z^i}\hfill \\ h(z)_{i>1}\frac{\eta _i}{z^i}+\underset{i>1}{}\frac{\eta _{ix}+\xi _i}{z^i}.\hfill \end{array}$$ (2.37) These equations give $`(a_i,h_i,\nu _i,\psi _i)`$ as differential polynomials in the $`(u_i,w_i,\xi _i,\eta _i)`$’s, and can be inverted modulo quadratures (as usual in the theory of KP–like equations). As we shall show in the next sections, there is some merit in considering such non standard coordinates, whose choice is suggested by the supersymmetric extension of Gel’fand–Zakharevich set-up for the KP theory. $`\mathrm{}`$ ### 2.5 A super KdV equation as a reduction of HSKP We will discuss now a supersymmetric generalization of the KdV equation obtained as reduction of the Hamiltonian super KP hierarchy. This example will be important in giving us one more clue to the second part of the paper. It can be shown that constraints of the form $$\widehat{H}^{(2k)}=z^k$$ are compatible with HSKP. We consider $`k=2`$ obtaining $$\begin{array}{c}\{\begin{array}{c}h_2=\frac{1}{2}h_1^{}+\psi _1\nu _1\hfill \\ h_k=\frac{1}{2}h_{k1}^{}\frac{1}{2}_{j=1}^{k2}h_{kj1}h_j+\psi _1\nu _{k1}\text{for }k>2\text{,}\hfill \end{array}\hfill \\ \{\begin{array}{c}\psi _2=\frac{1}{2}\psi _1^{}+a_1\psi _1\hfill \\ \psi _k=\frac{1}{2}\psi _{k1}^{}_{j=1}^{k2}h_{kj1}\psi _j+a_{k1}\psi _1\text{for }k>2\text{,}\hfill \end{array}\hfill \\ \{\begin{array}{c}\nu _2^{}=\frac{1}{2}\nu _1^{\prime \prime }+a_1^{}\nu _1\hfill \\ \nu _k^{}=\frac{1}{2}\nu _{k1}^{\prime \prime }_{j=1}^{k2}h_{kj1}\nu _j^{}+a_1^{}\nu _{k1}\text{for }k>2\text{,}\hfill \end{array}\hfill \\ \{\begin{array}{c}a_2^{}=\frac{1}{2}a_1^{\prime \prime }+a_1a_1^{}\hfill \\ a_k^{}=\frac{1}{2}a_{k1}^{\prime \prime }_{j=1}^{k2}h_{kj1}a_j^{}+a_{k1}a_1^{}\text{for }k>2.\hfill \end{array}\hfill \end{array}$$ (2.38) These equations allow us to compute recursively the coefficients $`h_j`$, $`\psi _j`$, $`\nu _j`$ and $`a_j`$ for $`j>1`$ in terms of $`h_1`$, $`\psi _1`$, $`\nu _1`$ and $`a_1`$ by means of quadratures. In order to explicitly write the equations for the independent degrees of freedom we have only to calculate the coefficients of order $`1`$ of the super current densities. Since the relations (2.38) algebraically determine only the derivatives of the fields $`a_i,\nu _i,ı2`$, in general the resulting equations will be integro–differential ones. Fortunately enough, for the sixth time of the hierarchy the non local terms cancel each other, and the result is $$\{\begin{array}{c}_6\nu _1=\frac{1}{4}\nu _1^{\prime \prime \prime }\frac{3}{2}a_1^{}\nu _1^{}3h_1\nu _1^{}\hfill \\ _6a_1=\frac{1}{4}a_1^{\prime \prime \prime }\frac{3}{2}a_{1}^{}{}_{}{}^{2}3h_1a_1^{}+6\psi _1\nu _1^{}\hfill \\ _6h_1=\frac{1}{4}h_1^{\prime \prime \prime }3h_1h_1^{}\frac{3}{2}\psi _1\nu _1^{\prime \prime }\frac{3}{2}\psi _1^{}\nu _1^{}\hfill \\ _6\psi _1=\frac{1}{4}\psi _1^{\prime \prime \prime }\frac{3}{2}a_1^{\prime \prime }\psi _1\frac{3}{2}a_1^{}\psi _1^{}3h_1\psi _1^{}3h_1^{}\psi _1\hfill \end{array}.$$ (2.39) We thus see that the evolution equations for the time $`t_6`$ are a supersymmetric extension of the KdV equation, which can be retrieved by setting $`a_1=\nu _1=\psi _1=0,h_1={\displaystyle \frac{u}{2}}`$. We notice also the following fact. Substituting $`\nu =\psi =0,h={\displaystyle \frac{u}{2}}`$ in the above equations (2.39) we obtain the ordinary system of PDE’s in two variables $`u`$ and $`a`$ $$\{\begin{array}{c}_ta=\frac{1}{4}a_{xxx}\frac{3}{2}a_{x}^{}{}_{}{}^{2}\frac{3}{2}ua_x\hfill \\ _tu=\frac{1}{4}u_{xxx}\frac{3}{2}uu_x\hfill \end{array}.$$ (2.40) One can easily notice that the submanifold defined by $`u=a_x+a^2+\lambda _0`$ is an invariant submanifold of these equations, where the first one coincides with the modified KdV equation. So we see that this reduction of HSKP “contains” both KdV and mKdV. This observation will be formalized and explained in the next sections. ## 3 HSKP and Darboux transformations In general, a Darboux transformation is a way to connect two systems of differential equations enabling to produce a solution of the second once a solution of the first has been supplied. This technique has proved to be very effective both in the construction of large classes of explicit solutions of soliton equations and in the understanding of the nature of infinite dimensional integrable systems (see, e.g., ). An example of such transformation is provided by the Miura map in the KdV theory and the modified KdV hierarchy (mKdV) (see, e.g., and the references quoted therein). Here we are mostly interested in the concept of Darboux intertwiners and Darboux coverings introduced in , where the geometrical features of the method where analyzed as follows. Consider three vector fields $`X`$, $`Y`$ and $`Z`$ on three manifolds $`M`$, $`N`$ and $`P`$, respectively. ###### Definition 3.1 The vector field $`Y`$ intertwines $`X`$ and $`Z`$ if there exists a pair of maps $`(\mu :NM,\sigma :NP)`$ such that $`X=\mu _{}Y`$ and $`Z=\sigma _{}Y`$. Moreover, if $`X=Z`$ , $`N`$ is a fiber bundle on $`M=P`$, and $`\mu :NM`$ is the bundle projection, then $`Y`$ is said to be a Darboux covering of $`X`$ , and the map $`\sigma `$ the associated Miura map. Finally, still when $`X=Z`$ and $`\mu :NM`$ is a fibre bundle, for each section $`\rho :MN`$ of $`\mu `$ which is invariant under $`Y`$, the composition $`\eta =\sigma \rho `$ which sends $`X`$ in $`X`$, and hence produces an integral curve $`\stackrel{~}{x}(t)`$ of $`X`$ from the integral curve $`x(t)`$ of $`X`$, is called a Darboux transformation. The concept of Darboux covering is useful for constructing both solutions and invariant submanifolds of the vector field $`X`$: if $`U`$ is a chart on $`M`$ with coordinate $`x`$ and $`V\mu ^1(U)`$ a chart on $`N`$ adapted to the projection $`\mu `$ and with fibered coordinates $`(x,a)`$, then the local expression of the above vector fields is $$\begin{array}{c}\dot{x}=X(x)\hfill \\ \dot{a}=Y(x,a),\hfill \end{array}$$ where the first equation is that of $`X`$ on $`U`$. Then, any integral curve $`x(t)`$ of $`X`$ can be lifted to an integral curve $`(x(t),a(t))`$ of $`Y`$ by solving the second equation, which is controlled by $`x(t)`$. Therefore, we get a new integral curve of $`X`$ by setting $$\stackrel{~}{x}(t)=\sigma (x(t),a(t)).$$ The last equation can also be interpreted as a “symmetry (or Darboux) transformation” of the dynamical system described by $`X`$, depending on a solution of the auxiliary system for $`a`$, controlled by $`X`$ itself, which associates $`\stackrel{~}{x}(t)`$ with $`x(t)`$. The application of the formalism we have just described to KP naturally leads to the DKP hierarchy. ###### Definition 3.2 Let $`M`$ be the affine space of (formal) monic Laurent series in $`z^1`$ with coefficients in $`C^{\mathrm{}}(S^1)`$ and of the form $$h(z,x)=z+\underset{j>0}{}h_j(x)z^j$$ and let $`N`$ be the affine space of couples $`(h,a)`$ where $`h`$ is as above and $`a`$ is a monic Laurent series of the form $$a(z,x)=z+\underset{j0}{}a_j(x)z^j.$$ Define two maps $`\mu ,\sigma :NM`$ by $$\mu (h,a)=h$$ and $$\stackrel{~}{h}:=\sigma (h,a)=h+\frac{_xa}{a}.$$ Finally, let $`H^{(k)}`$ and $`\stackrel{~}{H}^{(k)}`$ be the current densities associated with $`h`$ and $`\stackrel{~}{h}`$, respectively. The DKP hierarchy is the hierarchy of evolution equations on $`N`$ defined by $$\{\begin{array}{c}\frac{}{t_k}h=_xH^{(k)}\hfill \\ \frac{}{t_k}a=a(\stackrel{~}{H}^{(k)}H^{(k)})\hfill \end{array}.$$ DKP is a Darboux covering, in the sense of Definition 3.1, of the KP hierarchy $$_{t_k}h=_xH^{(k)}.$$ Indeed, it is clear that $`\mu _{}`$ maps the vector fields $`_{t_j}`$ of DKP to those of KP. As for $`\sigma _{}`$, we have $$\begin{array}{cc}\hfill _{t_k}\left(\frac{_xa}{a}\right)& =\frac{a_x_{t_k}a(_xa)(_{t_k}a)}{a^2}\hfill \\ \hfill =& \frac{a_x(a\stackrel{~}{H}^{(k)}aH^{(k)})a(_xa)(\stackrel{~}{H}^{(k)}H^{(k)})}{a^2}\hfill \\ \hfill =& _x\stackrel{~}{H}^{(k)}_xH^{(k)}\hfill \end{array}$$ (3.1) and finally $$_{t_k}\stackrel{~}{h}=_{t_k}h+_{t_k}\left(\frac{_xa}{a}\right)=_x\stackrel{~}{H}^{(k)}.$$ (3.2) In the papers , the following results were obtained: 1. The modified KP hierarchy of is the restriction of DKP on the invariant submanifold $`S_0N`$ defined by the simple equation $`a=h+a_0`$; 2. The DKP equations admit a remarkable family of invariant submanifolds, $`S_l`$, of which $`S_0`$ is the simplest; the images through $`\mu `$ of the intersections of two (or more) submanifolds, $`S_{l_1}\mathrm{}S_{l_k}`$ is an invariant submanifold of KP, which coincide with the rational KP reductions of Dickey and Krichever (see, e.g., ); 3. The central system CS can be explicitly linearized and classes of solutions can be explicitly found by means of a Darboux intertwiner linking it with the Sato system, that is, the coordinate expression of the linear flows of KP on the Sato Grassmannian. In the rest of this Section we will first show that HSKP (and hence JSKP) can be seen as a supersymmetric extension of DKP; then we will define a Darboux covering for HSKP, and briefly discuss the analogue of the invariant submanifolds mentioned in points 1 and 2 of the above list. The generalization of point 3 will be the subject of Section 4. ### 3.1 HSKP and Darboux transformations Our first goal is to give a connection between the Jacobian super KP hierarchy and DKP. First of all, we observe that the role $`a`$ has in DKP does not depend on the order of its pole, since it appears in a homogeneous way in all the equations. Hence, we see that (the reduction modulo nilpotents elements in $`\mathrm{\Lambda }`$ of) the bosonic degrees of freedom of HSKP are exactly the degrees of freedom of $`DKP`$: the Laurent series $`a`$ appearing in the definition (2.7) of the super Faà di Bruno generator can be identified with $`z^1`$ times the Laurent series $`a`$ appearing in the definition 3.2 of DKP. It is thus tempting to conjecture a relation between the two hierarchies. In fact, we can prove the following ###### Proposition 3.1 Let $`\widehat{h}`$ and $`\widehat{H}^{(k)}`$ be, respectively, the super Faà di Bruno generator and the currents of HSKP defined as in Section 2.1. * The constraint $`\nu =\psi =0`$ is compatible with the even flows of the HSKP hierarchy. * The reduction HSKP<sub>bos</sub> of the even flows of HSKP given by setting $`\nu =\psi =0`$ is isomorphic to DKP, i.e. if $`\widehat{h}`$ is a solution of HSKP<sub>bos</sub>, then $`(h,za)`$ is a solution of $`DKP`$ and vice versa. Proof. * Looking at the recurrence relations we introduced in Section 2.1, namely equations (2.9) and Table 1, we see that, under the constraint $`\nu =\psi =0`$, one has $$\{\begin{array}{c}\widehat{h}^{(2k1)}=\theta a^{(k)}+\phi h^{(k)}\hfill \\ \widehat{h}^{(2k)}=h^{(k)}(\theta \phi )b^{(k)},\hfill \end{array}$$ where the coefficients are given by the following recursion relations: $$\begin{array}{cc}& \{\begin{array}{c}h^{(k+1)}=(_x+h)h^{(k)}\hfill \\ h^{(0)}=1\hfill \end{array},\{\begin{array}{c}a^{(k+1)}=(_x+h)a^{(k)}\hfill \\ a^{(1)}=a\hfill \end{array},\hfill \\ & \{\begin{array}{c}b^{(k+1)}=(_x+h)b^{(k)}+(_xa)h^{(k)}\hfill \\ b^{(0)}=0.\hfill \end{array}\hfill \end{array}$$ (3.3) In particular $`\widehat{h}^{(0)}=1,\widehat{h}^{(1)}=\theta a+\phi h,\widehat{h}^{(2)}=h\theta \phi a_x.`$ This implies that $$\widehat{H}^{(2k)}=H^{(k)}(\theta \phi )K^{(k)},$$ where $`H^{(k)}`$ is the $`k`$-th current density of KP and $`K^{(k)}`$ is some power series in $`z^1`$ and $`x`$ of the form $$K^{(k)}(z,x)=\underset{j>0}{}K_j^k(x)z^j.$$ The evolution equations for the even flows of HSKP are then $$\{\begin{array}{c}_{t_{2k}}\nu =0\hfill \\ _{t_{2k}}\psi =0\hfill \\ _{t_{2k}}a=K^{(k)}\hfill \\ _{t_{2k}}h=_xH^{(k)},\hfill \end{array}$$ showing that the constraint $`\nu =\psi =0`$ is compatible with them. * In the proof of i. we have established also that $`h`$ evolves according to KP. We need only to understand better the evolution of $`a`$. We have to show that $`K^{(k)}/a+H^{(k)}`$ is the $`k`$-th current density of KP associated with $$\stackrel{~}{h}=h+\frac{_xa}{a}.$$ To achieve this we consider the function $$\widehat{A}=\left(\frac{1}{a}+\theta \phi \right),$$ and perform the “gauge transformation” $$\widehat{h}^{(k)}\widehat{l}^{(k)}:=\widehat{A}\widehat{h}^{(k)},$$ that is, we consider the new vector space $`\stackrel{~}{W}=\widehat{A}W`$ generated by the $`\widehat{l}^{(k)}`$’s. We remark that the first transformed basis elements are: $$\begin{array}{cc}& \widehat{l}^{(0)}=\frac{1}{a}+(\theta \phi )1,\widehat{l}^{(1)}=\theta \phi \frac{h}{a}\hfill \\ & \widehat{l}^{(2)}=\frac{h}{a}+(\theta \phi )\left(h+\frac{_xa}{a}\right)=\frac{h}{a}+(\theta \phi )\stackrel{~}{h}.\hfill \end{array}$$ Observe that since $`W`$ is generated by the action of the operator $`_x+\widehat{h}^{(2)}`$ on the pair $`(\widehat{h}^{(0)}=1,\widehat{h}^{(1)}=\theta a+\phi h)`$, $`\stackrel{~}{W}`$ will be generated by the action of $`\widehat{A}(_x+\widehat{h}^{(2)})\widehat{A}^1`$ on the pair $`(\widehat{l}^{(0)},\widehat{l}^{(1)})`$. We notice that $$\widehat{A}(_x+\widehat{h}^{(2)})\widehat{A}^1=_x+h+\frac{_xa}{a}=_x+\stackrel{~}{h},$$ which shows that $$\{\begin{array}{cc}& \widehat{l}^{(2k)}=f^{(k)}+\theta \phi \stackrel{~}{h}^{(k)}\hfill \\ & \widehat{l}^{(2k+1)}=\theta \stackrel{~}{h}^{(k)}+\phi g^{(k)},\hfill \end{array}$$ (3.4) Now we consider the transformed current $`\widehat{L}^{(2k)}=\widehat{A}\widehat{H}^{(2k)}`$. Its $`\theta \phi `$ component is clearly given by the sum $$\widehat{L}_{\theta \phi }^{(2k)}=H^{(k)}+K^{(k)}/a.$$ Since $`\widehat{L}^{(2k)}`$ is a finite linear combination of the basis elements $`\widehat{l}^{(j)}`$ it follows that $`\widehat{L}_{\theta \phi }^{(2k)}`$ is the unique combination of the $`\stackrel{~}{h}^{(k)}`$ with the asymptotics $$L_{\theta \phi }^{(2k)}=H^{(k)}+K^{(k)}/a=z^k+O(1/z).$$ Hence it must be equal to $`\stackrel{~}{H}^{(k)}`$, so we get the desired result $$K^{(k)}=a(H^{(k)}\stackrel{~}{H}^{(k)}).$$ $`\mathrm{}`$ ### 3.2 A Darboux covering for HSKP and the super analogue of its rational reductions In this section we return to the full supersymmetric picture, define the Darboux transformations and a D-HSKP hierarchy for the Hamiltonian super KP theory, and show how to obtain the super analogue of Dickey’s and Krichever’s rational reductions of the KP hierarchy. We observe that given a Laurent series $`\widehat{h}`$ of the usual form of equation (2.7), and a monic even power series $$\widehat{p}=p+\theta \zeta +\phi \xi +(\theta \phi )q,$$ with $`\overline{p}=\overline{q}=0`$, $`\overline{\zeta }=\overline{\xi }=1`$ and $$\{\begin{array}{c}p=1+_{j>0}p_jz^j\hfill \\ q=_{j>0}q_jz^j\hfill \\ \zeta =_{j>0}\zeta _jz^j\hfill \\ \xi =_{j>0}\xi _jz^j,\hfill \end{array}$$ the transformed series $$\widehat{k}=\widehat{h}+\frac{\delta \widehat{p}}{\widehat{p}}.$$ is still of type (2.7). ###### Definition 3.3 (D-HSKP) Let $`\widehat{N}`$ be the affine space of couples of monic formal Laurent series $`(\widehat{h},\widehat{p})`$, let $$\widehat{k}=\widehat{h}+\frac{\delta \widehat{p}}{\widehat{p}}$$ and let $`\widehat{K}^{(k)}`$ be the $`k`$-th super current density associated to $`\widehat{k}`$. The Darboux–Hamiltonian super KP hierarchy is the set of compatible evolution equations $$\{\begin{array}{c}_{t_k}\widehat{h}=(1)^k\delta \widehat{H}^{(k)}\hfill \\ _{t_k}\widehat{p}=\widehat{p}(\widehat{K}^{(k)}\widehat{H}^{(k)})\hfill \end{array}.$$ If we let $`\widehat{M}`$ be the affine space of the monic formal Laurent series $`\widehat{h}`$ and we define two maps $`\widehat{\mu },\widehat{\sigma }:\widehat{N}\widehat{M}`$ by $$\widehat{\mu }(\widehat{h},\widehat{p})=\widehat{h}$$ and $$\widehat{\sigma }(\widehat{h},\widehat{p})=\widehat{h}+\frac{\delta \widehat{p}}{\widehat{p}},$$ then $$\begin{array}{cc}\hfill _{t_k}\left(\frac{\delta \widehat{p}}{\widehat{p}}\right)& =(1)^k\frac{\widehat{p}\delta _{t_k}\widehat{p}(\delta \widehat{p})(_{t_k}\widehat{p})}{\widehat{p}^2}\hfill \\ \hfill =& (1)^k\frac{\widehat{p}\delta (\widehat{p}\widehat{K}^{(k)}\widehat{p}\widehat{H}^{(k)})\widehat{p}(\delta \widehat{p})(\widehat{K}^{(k)}\widehat{H}^{(k)})}{\widehat{p}^2}\hfill \\ \hfill =& (1)^k\delta (\widehat{K}^{(k)}\widehat{H}^{(k)})\hfill \end{array}$$ so $$_{t_k}\widehat{k}=_{t_k}\widehat{h}+_{t_k}\left(\frac{\delta \widehat{p}}{\widehat{p}}\right)=(1)^k\delta \widehat{K}^{(k)},$$ i.e. D-HSKP is a Darboux covering of HSKP. In the next section we will use this formalism for a geometrical characterization of the analogue of the rational reductions of the KP hierarchy. ### 3.3 Super mKP and Rational Hierarchies ###### Proposition 3.2 The submanifold $`\widehat{𝒮}_l`$ of $`\widehat{N}`$ (see Definition 3.3) characterized by $$z^{l/2}\widehat{p}W$$ for $`l`$ even, or by $$\theta z^{(l1)/2}\widehat{p}W$$ for $`l`$ odd, is invariant under the flows of the D-HSKP hierarchy, where we recall that $`W=\mathrm{span}_{B_{[x\phi ]}}\{\widehat{h}^{(j)}|j0\}.`$ Consequently, the submanifold $$\widehat{𝒯}_l:=\widehat{\mu }(\widehat{𝒮}_l)$$ of $`\widehat{M}`$ is invariant under HSKP. Proof. We give the proof only for $`l=2n`$ even, the other proof is the same up to some obvious change of signs. The condition $`(\widehat{h},\widehat{p})\widehat{𝒮}_l`$ implies the existence of some coefficients $`\alpha _j(x,\phi )`$, $`j=0,\mathrm{},l`$ such that $$z^n\widehat{p}=\underset{j=0}{\overset{l}{}}\alpha _j\widehat{H}^{(j)},$$ (3.5) so we have to show that this expression is invariant under the flows of D-HSKP, i.e. $$_{t_k}\left(z^l\widehat{p}\underset{j=0}{\overset{l}{}}\alpha _j\widehat{H}^{(j)}\right)=0$$ (3.6) on $`\widehat{𝒮}_{2n}`$. Let $`W^{\widehat{k}}:=\mathrm{span}_{B_{[x\phi ]}}\{\widehat{k}^{(j)}|j0\}`$. By definition we have $$\widehat{p}(\delta +\widehat{k})=(\delta +\widehat{h})\widehat{p},$$ and hence $`\widehat{p}(\delta +\widehat{k})^j=(\delta +\widehat{h})^j\widehat{p}.`$ This implies that $`z^l\widehat{p}W^{\widehat{k}}W,`$ and, therefore, using the D-JSKP equations, $`(_{t_k}+\widehat{H}^{(k)})z^l\widehat{p}=z^l\widehat{p}\widehat{K}^{(k)}W,`$ i.e. $$z^l_{t_k}\widehat{p}+\underset{j=0}{\overset{l}{}}(1)^{jk}\alpha _j\widehat{H}^{(k)}\widehat{H}^{(j)}W.$$ Using now the property $`_{t_k}\widehat{H}^{(j)}+\widehat{H}^{(k)}\widehat{H}^{(j)}W`$, characteristic of HSKP, and comparing the coefficients of $`z^j`$ and $`\theta z^j`$ in equation (3.5) for $`j=0,\mathrm{},l`$, we get $$z^l_{t_k}\widehat{p}\underset{j=0}{\overset{l}{}}(1)^{jk}\alpha _j_{t_k}\widehat{H}^{(j)}=\underset{j=0}{\overset{l}{}}(_{t_k}\alpha _j)\widehat{H}^{(j)},$$ i.e. (3.6) holds. $`\mathrm{}`$ As a first application of this result we show how, in such a formalism, we obtain a supersymmetric extension of the modified KP hierarchy. We consider the submanifold $`\widehat{𝒮}_2`$ defined by $`z\widehat{p}W`$. It can be seen that these equations entail the following constraints: $$\zeta _1=0;_{S^1}\xi _j𝑑x=0,j2;q_1=0;\text{ and }_{S^1}q_i𝑑x=0j2.$$ (3.7) The bosonic sector of the resulting theory covers the invariant submanifold $`S_0`$ of the DKP equations of defined by $`a=h+a_0`$. There it was proven that $`DKP|_{S_0}`$ is another form of the modified KP theory of Kupershmidt . Hence, through this result, we obtain that the restriction D–SKP$`_{|_{𝒮_2}}`$ provides a direct supersymmetric extension of mKP. Finally, following , one can define and study the “rational–type reductions” of the HSKP hierarchy as the restriction of the D-HSKP hierarchy to the intersection of suitable of invariant submanifolds. In the next example we will briefly describe the simplest case. ###### Example 3.1 Let us consider the triple intersection $`\widehat{𝒮}_{124}`$, and its image $`\widehat{𝒯}_{124}`$ under $`\widehat{\mu }`$, obtained by requiring that the triple $`(\theta \widehat{p},z\widehat{p},z^2\widehat{p})`$ lie in $`W`$. Since $`\widehat{H}^{(1)}=\theta a+\nu `$ we see that, recalling the form of $`\widehat{p}=p+\theta \zeta +\phi \xi +(\theta \phi )q`$, the equation $`\theta \widehat{p}W`$ implies $$\nu =0,\xi =0,a=p.$$ Intersecting with $`\widehat{𝒮}_2`$ we get $$h=zpp_1,\psi =z\zeta +\zeta _1p,q=\frac{q_1pp_x}{z}.$$ Finally, requiring $`z^2\widehat{p}W`$ one sees that it is possible to express all the fields $`p_i,q_i,\zeta _i,`$ (and hence all the currents $`\widehat{H}^{(j)}`$) in terms of the two even fields $`r=p_1,s=p_2`$ and the two odd fields $`\rho =\zeta _1,\sigma =\zeta _2`$. Indeed the equations to be solved are: $$\{\begin{array}{c}z^2(p^2p)+z(p_xp_1p)=p_{1,x}+p_2;\hfill \\ z^2(2p\zeta \zeta )+z(\zeta _x\zeta _1p^2p_1\zeta )=\zeta _1p+\zeta _2+p_1p\zeta ,\hfill \\ q_1\left(z(p^2p)+p_xp_1p\right)=z(2pp_xp_x)+p_{xx}p_{1x}pp_1p_x.\hfill \end{array}$$ From the first one we get $$p_{k+2}=p_{k+1x}\underset{j=1}{\overset{k}{}}p_jp_{kj+2},k1$$ and from the second a similar formula expressing $`\zeta _{j+2},j1`$ in terms of $`\zeta _1,\zeta _2`$ and the $`p_k`$’s. Plugging the first equation into the third one, we finally get $$q_1=\frac{d}{dx}\mathrm{log}(p_2+p_{1x}).$$ The resulting equations of motion relative to the time $`t_4`$ are the following: $$\{\begin{array}{cc}\hfill \dot{r}& =\left(r_x+2sr^2\right)_x;\hfill \\ \hfill \dot{s}& =\left(s_x+rs\right)_x;\hfill \\ \hfill \dot{\rho }& =\rho _{xx}+2\sigma _x2r\rho _x2r_x\rho +2\left(\mathrm{log}(r_x+s)\right)_x\left(\rho _x+\sigma r\rho \right)\hfill \\ \hfill \dot{\sigma }& =\sigma _{xx}+2r\rho _{xx}2(r^2+s)\rho _x2r_x\sigma +2r\left(\mathrm{log}(r_x+s)\right)_x\left(\rho _x+\sigma r\rho \right).\hfill \end{array}$$ (3.8) We notice that these evolution equations for $`r`$ and $`s`$ coincide with those of the realization of the well-known AKNS (or two–boson) hierarchy as a rational reduction of the KP hierarchy . $`\mathrm{}`$ ## 4 Linearization The evolution equations of SCS we have introduced in Section 2.2 are not linear, and not directly linearizable. To obtain their “linearized version”, allowing to provide explicit solutions, we can exploit Darboux covering techniques as it has been done in for KP. The idea is to find a Darboux covering which intertwines the super Central System SCS defined in Section 2.1 with a new hierarchy whose linearization can be achieved by elementary methods. To this end, let $`\widehat{}`$ be the space of sequences of Laurent series $`\{\widehat{Y}^{(k)}\}_{k0}`$ of the form $$\{\begin{array}{c}\widehat{Y}^{(2k)}=z^k+_{j>0}\left(\widehat{Y}_{0,j}^{2k}z^j+\widehat{Y}_{1,j}^{2k}\theta z^j\right)\hfill \\ \widehat{Y}^{(2k+1)}=\theta z^k+_{j>0}\left(\widehat{Y}_{0,j}^{2k+1}z^j+\widehat{Y}_{1,j}^{2k+1}\theta z^j\right),\hfill \end{array}$$ where $`\overline{\widehat{Y}^{(k)}}=kmod2`$. The third manifold $`\widehat{𝒫}`$ of Definition 3.1 is just a copy of $`\widehat{}`$ formed by the sequences $`\{\widehat{H}^{(k)}\}_{k0}`$. Finally, the manifold $`\widehat{𝒩}`$ is the Cartesian product $`\widehat{}\times \widehat{𝒢}`$ of $`\widehat{}`$ by the group of even invertible formal power series $`\widehat{w}`$ of the form $$\widehat{w}=1+\underset{j>0}{}\left(\widehat{w}_{0,j}z^j+\widehat{w}_{1,j}\theta z^j\right).$$ The next step is to define suitable vector fields $`\widehat{𝒳}`$, $`\widehat{𝒴}`$ and $`\widehat{𝒵}`$ on $`\widehat{}`$, $`\widehat{𝒩}`$ and $`\widehat{𝒫}`$, respectively. The vector field $`\widehat{𝒵}`$ is any vector field of SCS, which is completely characterized by $$(_{t_k}+\widehat{H}^{(k)})WW.$$ The flow can be identified by using an index, so we call this vector field $`\widehat{𝒵}_k`$. To define $`\widehat{𝒳}`$ we introduce the subspace $`W^{(\widehat{Y})}`$ of $`V`$ spanned by the $`\widehat{Y}^{(j)}`$’s. Then, if $`k=2n`$ is even we let $`\widehat{𝒳}_k`$ be the vector field characterized by the property $$(_{t_k}+z^n)W^{(\widehat{Y})}W^{(\widehat{Y})},$$ while if $`k=2n+1`$ we let $`\widehat{𝒳}_k`$ be the vector field characterized by $$(_{t_k}+\theta z^n)W^{(\widehat{Y})}W^{(\widehat{Y})}.$$ As for SCS, we can write down the equations defining $`\widehat{𝒳}_k`$ by comparing coefficients: if $`k=2n`$ $$_{t_k}\widehat{Y}^{(j)}+z^n\widehat{Y}^{(j)}=\widehat{Y}^{(j+2n)}+\underset{l=1}{\overset{n}{}}\left(\widehat{Y}_{0,l}^j\widehat{Y}^{(2n2l)}+\widehat{Y}_{1,l}^j\widehat{Y}^{(2n2l+1)}\right),$$ while if $`k=2n+1`$ $$\{\begin{array}{c}_{t_k}\widehat{Y}^{(2j)}+\theta z^n\widehat{Y}^{(2j)}=\widehat{Y}^{(2j+2n+1)}+_{l=1}^n\widehat{Y}_{0,l}^{2j}\widehat{Y}^{(2n2l+1)}\hfill \\ _{t_k}\widehat{Y}^{(2j+1)}+\theta z^n\widehat{Y}^{(2j+1)}=_{l=1}^n\widehat{Y}_{0,l}^{2j+1}\widehat{Y}^{(2n2l+1)}.\hfill \end{array}$$ The definition of $`\widehat{𝒴}_k`$ is obtained imposing the further condition $$(_{t_k}+z^n)\widehat{w}W^{(\widehat{Y})}\text{ if }k=2n,$$ $$(_{t_k}+\theta z^n)\widehat{w}W^{(\widehat{Y})}\text{ if }k=2n+1.$$ As in , we give the following ###### Definition 4.1 We call super Sato System (SS) the family of vector fields $`\{\widehat{𝒳}_k\}_{k>0}`$ on $`\widehat{}`$ and super Darboux–Sato System (SDS) the family of vector fields $`\{\widehat{𝒴}_k\}_{k>0}`$ on $`\widehat{𝒩}`$. The next step is to define the maps $`\widehat{\mu }:\widehat{𝒩}\widehat{}`$ and $`\widehat{\sigma }:\widehat{𝒩}\widehat{𝒫}`$. The first is as usual the projection $$(\{\widehat{Y}^{(k)}\}_{k0},\widehat{w})\{\widehat{Y}^{(k)}\}_{k0},$$ while the second is defined by imposing the intertwining condition $$\widehat{w}W=W^{(\widehat{Y})},$$ which holds if and only if $$\{\begin{array}{c}\widehat{w}\widehat{H}^{(2j)}=\widehat{Y}^{(2j)}+_{l=1}^j\left(\widehat{w}_{0,l}\widehat{Y}^{(2j2l)}+\widehat{w}_{1,l}\widehat{Y}^{(2j2l+1)}\right)\hfill \\ \widehat{w}\widehat{H}^{(2j+1)}=\widehat{Y}^{(2j+1)}+_{l=1}^j\widehat{w}_{0,l}\widehat{Y}^{(2j2l+1)}.\hfill \end{array}$$ ###### Lemma 4.1 The SDS system is a Darboux intertwiner of SS with SCS. Proof. We need only to prove that $`\widehat{\sigma }_{}(SDS)=SCS`$. This follows by observing that the definitions of SDS and $`\widehat{\sigma }`$ imply $$\{\begin{array}{c}_{t_{2k}}\widehat{w}+z^k\widehat{w}=\widehat{w}\widehat{H}^{(2k)}\hfill \\ _{t_{2k+1}}\widehat{w}+\theta z^k\widehat{w}=\widehat{H}^{(2k+1)},\hfill \end{array}$$ so $$\{\begin{array}{c}\widehat{w}(_{t_{2k}}+\widehat{H}^{(2k)})=(_{t_{2k}}+z^k)\widehat{w}\hfill \\ \widehat{w}(_{t_{2k+1}}+\widehat{H}^{(2k+1)})=(_{t_{2k}}+\theta z^k)\widehat{w}.\hfill \end{array}$$ Hence, we get $`\widehat{w}(_{t_{2k}}+\widehat{H}^{(2k)})W`$ $`=`$ $`(_{t_{2k}}+z^k)\widehat{w}W`$ $`=`$ $`(_{t_{2k}}+z^k)W^{(\widehat{Y})}W^{(\widehat{Y})}`$ and $`\widehat{w}(_{t_{2k+1}}+\widehat{H}^{(2k+1)})W`$ $`=`$ $`(_{t_{2k+1}}+\theta z^k)\widehat{w}W`$ $`=`$ $`(_{t_{2k+1}}+\theta z^k)W^{(\widehat{Y})}W^{(\widehat{Y})},`$ showing that $`(_{t_k}+\widehat{H}^{(k)})WW`$, i.e. the SCS. $`\mathrm{}`$ We consider now the map $`\widehat{\rho }:\widehat{}\widehat{𝒩}`$ defined by $$\{\widehat{Y}^{(k)}\}_{k0}(\{\widehat{Y}^{(k)}\}_{k0},\widehat{Y}^{(0)})$$ and the corresponding map $`\widehat{\sigma }\widehat{\rho }:\widehat{}\widehat{𝒫}`$. ###### Lemma 4.2 The submanifold $`\widehat{\rho }(\widehat{})`$ of $`\widehat{𝒩}`$ is a section of $`\widehat{\mu }`$ invariant under SDS. Proof. The previous definitions imply $$\{\begin{array}{c}_{2j}(\widehat{w}\widehat{Y}^{(0)})=z^j(\widehat{w}\widehat{Y}^{(0)})\hfill \\ +_{l=1}^j\left((\widehat{w}_{0,l}\widehat{Y}_{0,l}^0)\widehat{Y}^{(2j2l)}+(\widehat{w}_{1,l}\widehat{Y}_{1,l}^0)\widehat{Y}^{(2j2l+1)}\right)\hfill \\ \\ _{2j+1}(\widehat{w}\widehat{Y}^{(0)})=\theta z^j(\widehat{w}\widehat{Y}^{(0)})+_{l=1}^j(\widehat{w}_{0,l}\widehat{Y}_{0,l}^0))\widehat{Y}^{(2j2l+1)}\hfill \end{array},$$ proving the lemma. $`\mathrm{}`$ We have now to linearize the super Sato system. To achieve the result it is better to introduce the infinite even matrix $`𝖸`$ defined by $$𝖸_{jk}:=\{\begin{array}{cc}\widehat{Y}_{0,\frac{k+2}{2}}^j\hfill & \text{for }k\text{ even}\hfill \\ & \\ \widehat{Y}_{1,\frac{k+1}{2}}^j\hfill & \text{for }k\text{ odd},\hfill \end{array}$$ where $`j,k0`$, and the associated matrix $`\stackrel{~}{𝖸}`$ whose entries are $$\stackrel{~}{𝖸}_{jk}:=(1)^{\overline{𝖸}_{jk}}𝖸_{jk}=(1)^{j+k}𝖸_{jk}.$$ An easy computation shows that the flows of the SS hierarchy translate into the following Riccati type evolution equations: $$\{\begin{array}{c}_{2n}𝖸+𝖸\mathrm{\Lambda }_{2}^{t}{}_{}{}^{n}\mathrm{\Lambda }_2^n𝖸=𝖸\mathrm{\Gamma }_{2n}𝖸\hfill \\ _{2n+1}𝖸+\stackrel{~}{𝖸}\mathrm{\Lambda }_1\mathrm{\Lambda }_{2}^{t}{}_{}{}^{n}\mathrm{\Lambda }_1\mathrm{\Lambda }_2^n𝖸=\stackrel{~}{𝖸}\mathrm{\Gamma }_{2n+1}𝖸\hfill \end{array},$$ (4.1) where <sup>t</sup> means ordinary transposition (not super transposition), $`\mathrm{\Lambda }_1`$ is the odd shift matrix with entries $$(\mathrm{\Lambda }_1)_{jk}:=\frac{1(1)^k}{2}\delta _{k,j+1},$$ $`\mathrm{\Lambda }_2`$ is the even shift matrix with entries $$(\mathrm{\Lambda }_2)_{jk}:=\delta _{k,j+2},$$ $`\mathrm{\Gamma }_{2n}`$ is the even convolution matrix defined by $$(\mathrm{\Gamma }_{2n})_{jk}:=\frac{1(1)^k}{2}\delta _{k,2nj}+\frac{1(1)^{k+1}}{2}\delta _{k,2nj2}$$ and finally $`\mathrm{\Gamma }_{2n+1}`$ is the odd convolution matrix given by $$(\mathrm{\Gamma }_{2n+1})_{jk}:=\frac{1(1)^k}{2}\delta _{k,2nj1}.$$ Observe that these matrices satisfy the relations $$[\mathrm{\Lambda }_1,\mathrm{\Lambda }_1]=[\mathrm{\Lambda }_1,\mathrm{\Lambda }_2]=[\mathrm{\Lambda }_1,\mathrm{\Lambda }_2^t]=0,$$ $$\mathrm{\Lambda }_2^t\mathrm{\Gamma }_n=\mathrm{\Gamma }_n\mathrm{\Lambda }_2,\mathrm{\Lambda }_1\mathrm{\Gamma }_{2n}=\mathrm{\Gamma }_{2n}\mathrm{\Lambda }_1,\text{ and }\mathrm{\Lambda }_1\mathrm{\Gamma }_{2n+1}=\mathrm{\Gamma }_{2n+1}\mathrm{\Lambda }_1=0,$$ which imply the compatibility of the above system of matrix Riccati equations. ###### Proposition 4.3 The infinite even matrix $`𝖸`$ is a solution of (4.1) if and only if it has the form $`𝖸=𝖵𝖴^1`$, where $`𝖴`$ and $`𝖵`$ are infinite even matrices satisfying the constant coefficients linear system $$\{\begin{array}{c}_{2n}𝖴=\mathrm{\Lambda }_{2}^{t}{}_{}{}^{n}𝖴\mathrm{\Gamma }_{2n}𝖵\hfill \\ _{2n+1}𝖴=\mathrm{\Lambda }_1\mathrm{\Lambda }_{2}^{t}{}_{}{}^{n}𝖴\mathrm{\Gamma }_{2n+1}𝖵\hfill \\ _{2n}𝖵=\mathrm{\Lambda }_2^n𝖵\hfill \\ _{2n+1}𝖵=\mathrm{\Lambda }_1\mathrm{\Lambda }_2^n𝖵\hfill \end{array}$$ with, of course, $`𝖴`$ invertible. Proof. The proof is exactly the same as in the commutative case, once we have observed that for two matrices $`𝖴`$ and $`𝖵`$ the following relations hold: $$\{\begin{array}{c}_{t_{2k}}(\mathrm{𝖴𝖵})=(_{t_{2k}}𝖴)𝖵+𝖴(_{t_{2k}}𝖵)\hfill \\ _{t_{2k+1}}(\mathrm{𝖴𝖵})=(_{t_{2k+1}}𝖴)𝖵+\stackrel{~}{𝖴}(_{t_{2k+1}}𝖵)\hfill \\ \stackrel{~}{\mathrm{𝖴𝖵}}=\stackrel{~}{𝖴}\stackrel{~}{𝖵}\stackrel{~}{𝖴^1}=\stackrel{~}{𝖴}^1\hfill \end{array}.$$ Thus, if $`𝖴`$ and $`𝖵`$ solve the system of linear equations of the statement and if we let $`𝖸=\mathrm{𝖵𝖴}^1`$, then $`_{2n}𝖸`$ $`=`$ $`(_{2n}𝖵)𝖴^1\mathrm{𝖵𝖴}^1(_{2n}𝖴)𝖴^1`$ $`=`$ $`\mathrm{\Lambda }_2^n\mathrm{𝖵𝖴}^1\mathrm{𝖵𝖴}^1\mathrm{\Lambda }_{2}^{t}{}_{}{}^{n}+\mathrm{𝖵𝖴}^1\mathrm{\Gamma }_{2n}\mathrm{𝖵𝖴}^1`$ $`=`$ $`𝖸\mathrm{\Lambda }_{2}^{t}{}_{}{}^{n}+\mathrm{\Lambda }_2^n𝖸+𝖸\mathrm{\Gamma }_{2n}𝖸`$ and $`_{2n+1}𝖸`$ $`=`$ $`(_{2n+1}𝖵)𝖴^1\stackrel{~}{𝖵}\stackrel{~}{𝖴}^1(_{2n+1}𝖴)𝖴^1`$ $`=`$ $`\mathrm{\Lambda }_1\mathrm{\Lambda }_2^n\mathrm{𝖵𝖴}^1\stackrel{~}{𝖵}\stackrel{~}{𝖴}^1\mathrm{\Lambda }_1\mathrm{\Lambda }_{2}^{t}{}_{}{}^{n}+\stackrel{~}{𝖵}\stackrel{~}{𝖴}^1\mathrm{\Gamma }_{2n+1}\mathrm{𝖵𝖴}^1`$ $`=`$ $`\stackrel{~}{𝖸}\mathrm{\Lambda }_1\mathrm{\Lambda }_{2}^{t}{}_{}{}^{n}+\mathrm{\Lambda }_1\mathrm{\Lambda }_2^n𝖸+\stackrel{~}{𝖸}\mathrm{\Gamma }_{2n}𝖸.`$ Therefore, if we look for a solution $`𝖸`$ of the Riccati matrix equations of SS with initial condition $`𝖸(0)=𝖸_0`$, we have simply to solve the linear system above imposing the initial conditions $`𝖵(0)=𝖸_0`$ and $`𝖴(0)=𝕀`$. As we already noticed, the necessary condition $$\mathrm{\Gamma }_{2n+1}\mathrm{\Lambda }_1=0$$ for the integrability of the linear system holds. $`\mathrm{}`$ Of course, the computations given in the proposition are only formal: to make sense of them one should also introduce a suitable notion of convergence for the intervening series in infinite variables. However, notice that the constraint “$`𝖸_{jk}=0`$ when either $`jJ`$ or $`kK`$” is compatible with the evolution equations for $`𝖸`$, allowing us to consider reductions where only the finite submatrix $`𝖸_{JK}`$ of $`𝖸`$ consisting of its first $`J`$ rows and $`K`$ columns does not vanish. Obviously, $`𝖸_{JK}`$ evolves according to the reduced Riccati equations $$\{\begin{array}{c}_{2n}𝖸_{JK}+𝖸_{JK}\mathrm{\Lambda }_{2,KK}^{t}{}_{}{}^{n}\mathrm{\Lambda }_{2,JJ}^n𝖸_{JK}=𝖸_{JK}\mathrm{\Gamma }_{2n,KJ}𝖸_{JK}\hfill \\ _{2n+1}𝖸_{JK}+\stackrel{~}{𝖸}_{JK}(\mathrm{\Lambda }_1\mathrm{\Lambda }_{2}^{t}{}_{}{}^{n})_{KK}\mathrm{\Lambda }_{1,JJ}\mathrm{\Lambda }_{2,JJ}^n𝖸_{JK}=\stackrel{~}{𝖸}_{JK}\mathrm{\Gamma }_{2n+1,KJ}𝖸_{JK}\hfill \end{array}.$$ This is a closed system of (graded) ordinary differential equations in a finite number of variables. It yields “finite type” solutions (i.e. depending only on finitely many times) of SS and hence of SCS and HSKP. Observe that the compatibility of the reduced system requires $`K`$ to be even; in this case $`(\mathrm{\Lambda }_1\mathrm{\Lambda }_{2}^{t}{}_{}{}^{n})_{KK}=\mathrm{\Lambda }_{1,KK}\mathrm{\Lambda }_{2,KK}^{t}{}_{}{}^{n}`$. ### 4.1 An explicit example To show an example, we compute the solution of SS associated to $`J=3`$ and $`K=4`$. To simplify notations let us call $`Y:=𝖸_{34}`$, $$A_1:=\mathrm{\Lambda }_{1,33}=\left(\begin{array}{ccc}0& 1& 0\\ 0& 0& 0\\ 0& 0& 0\end{array}\right),A_2:=\mathrm{\Lambda }_{2,33}=\left(\begin{array}{ccc}0& 0& 1\\ 0& 0& 0\\ 0& 0& 0\end{array}\right),$$ $$B_1:=\mathrm{\Lambda }_{1,44}=\left(\begin{array}{cccc}0& 1& 0& 0\\ 0& 0& 0& 0\\ 0& 0& 0& 1\\ 0& 0& 0& 0\end{array}\right),B_2:=\mathrm{\Lambda }_{2,44}^t=\left(\begin{array}{cccc}0& 0& 0& 0\\ 0& 0& 0& 0\\ 1& 0& 0& 0\\ 0& 1& 0& 0\end{array}\right),$$ and $`C_k:=\mathrm{\Gamma }_{k,43}`$. The relevant (i.e. different from zero) convolution matrices are $$C_2:=\left(\begin{array}{ccc}1& 0& 0\\ 0& 1& 0\\ 0& 0& 0\\ 0& 0& 0\end{array}\right),C_3:=\left(\begin{array}{ccc}0& 1& 0\\ 0& 0& 0\\ 0& 0& 0\\ 0& 0& 0\end{array}\right),$$ $$C_4:=\left(\begin{array}{ccc}0& 0& 1\\ 0& 0& 0\\ 1& 0& 0\\ 0& 1& 0\end{array}\right),C_5:=\left(\begin{array}{ccc}0& 0& 0\\ 0& 0& 0\\ 0& 1& 0\\ 0& 0& 0\end{array}\right),C_6:=\left(\begin{array}{ccc}0& 0& 0\\ 0& 0& 0\\ 0& 0& 1\\ 0& 0& 0\end{array}\right).$$ We see that $`A_2^2=0`$ and $`B_2^2=0`$, so the solution of SS (or SCS) will depend only on the first six times. We solve the Riccati system for $`Y`$ by introducing the $`4\times 4`$ matrix $`U`$ and the $`3\times 4`$ matrix $`V`$ which are solutions of the following linear Cauchy problems: $$\{\begin{array}{c}_{t_{2k}}V=A_2^kV\hfill \\ _{t_{2k+1}}V=A_1A_2^kV\hfill \\ \\ V(0)=Y(0)\hfill \end{array}\{\begin{array}{c}_{t_{2k}}U=B_2^kUC_{2k}V\hfill \\ _{t_{2k+1}}U=B_1B_2^kUC_{2k+1}V\hfill \\ \\ U(0)=𝕀\hfill \end{array}$$ and then putting $`Y:=VU^1`$. First of all we find that $$V=\mathrm{exp}\underset{j>0}{}(t_{2j}A_2^j+t_{2j1}A_1A_2^{j1})V(0)=\left(\begin{array}{ccc}1& t_1& t_2\\ 0& 1& 0\\ 0& 0& 1\end{array}\right)Y(0).$$ Then we solve the system for $`U`$ by introducing the matrix $`U_0`$ defined by $$U=\mathrm{exp}\underset{j>0}{}(t_{2j}B_2^j+t_{2j1}B_1B_2^{j1})U_0=(𝕀+t_1B_1+t_2B_2+(t_3+t_1t_2)B_1B_2)U_0$$ and evolving as $$\{\begin{array}{c}_{t_{2k}}U_0=(𝕀t_1B_1t_2B_2(t_3t_1t_2)B_1B_2)C_{2k}V\hfill \\ _{t_{2k+1}}U_0=(𝕀+t_1B_1t_2B_2+(t_3t_1t_2)B_1B_2)C_{2k+1}V\hfill \end{array}.$$ The equations for $`U_0`$ are easily solvable and we get $$U_0=𝕀\left(\begin{array}{ccc}t_2& t_3& t_4+\frac{1}{2}t_2^2\\ 0& t_2& 0\\ t_4\frac{1}{2}t_2^2& t_5t_2t_3& t_6\frac{1}{3}t_2^3\\ 0& t_4\frac{1}{2}t_2^2& 0\end{array}\right)Y(0).$$ In order to write down an effective solution, we choose simple initial conditions, e.g. $$Y(0)=\left(\begin{array}{cccc}0& 0& 0& 0\\ 0& 0& 0& 0\\ 0& 0& 1& 0\end{array}\right).$$ Then $$V=\left(\begin{array}{cccc}0& 0& t_2& 0\\ 0& 0& 0& 0\\ 0& 0& 1& 0\end{array}\right),$$ $$U=\left(\begin{array}{cccc}1& t_1& t_4\frac{1}{2}t_2^2& 0\\ 0& 1& 0& 0\\ t_2& t_3+t_1t_2& 1t_6t_2t_4\frac{1}{6}t_2^3& t_1\\ 0& t_2& 0& 1\end{array}\right).$$ Finally, we find $$\widehat{Y}^{(0)}=1\frac{3t_2^2}{\tau }z^1+\frac{3t_2(t_1t_2t_3)}{\tau }\theta z^1+\frac{3t_2}{\tau }z^2\frac{3t_1t_2}{\tau }\theta z^2,$$ $$\widehat{Y}^{(1)}=\theta ,$$ $$\widehat{Y}^{(2)}=z\frac{3t_2}{\tau }z^1+\frac{3(t_1t_2t_3)}{\tau }\theta z^1+\frac{3}{\tau }z^2\frac{3t_1}{\tau }\theta z^2,$$ $$\widehat{Y}^{(2k)}=z^k\text{for }k>1$$ $$\widehat{Y}^{(2k+1)}=\theta z^k\text{for }k>0,$$ where $`\tau =3+t_2^33t_6`$. We can thus compute the first super currents of SCS $$\widehat{H}^{(1)}=\theta +\underset{k>0}{}\left(\frac{3t_2^2}{\tau }z^1+\frac{3t_2}{\tau }z^2\right)^k\theta ,$$ $`\widehat{H}^{(2)}`$ $`=`$ $`z{\displaystyle \frac{3t_2^2}{\tau }}+3{\displaystyle \underset{k0}{}}({\displaystyle \frac{3t_2^2}{\tau }}z^1{\displaystyle \frac{3t_2}{\tau }}z^2)^k\times `$ $`\left({\displaystyle \frac{t_2^2}{\tau }}{\displaystyle \frac{2t_2}{\tau }}z^1+{\displaystyle \frac{2t_1t_2t_3}{\tau }}\theta z^1+{\displaystyle \frac{1}{\tau }}z^2{\displaystyle \frac{t_1}{\tau }}\theta z^2\right).`$ As explained in Section 2.2, we obtain a solution of HSKP after substituting $`t_2`$ and $`t_1`$ with $`x`$ and $`\phi +t_1`$ respectively and putting $`\widehat{h}=\widehat{H}^{(1)}+\phi \widehat{H}^{(2)}`$: $$\{\begin{array}{c}\nu =0\hfill \\ a=1+_{k>0}\left(\frac{3x^2}{\tau }z^1\frac{3x}{\tau }z^2\right)^k\hfill \\ h=z\frac{6x}{\tau }z^1+\frac{3}{\tau }z^2+3_{k>0}\left(\frac{3x^2}{\tau }z^1\frac{3x}{\tau }z^2\right)^k\left(\frac{x^2}{\tau }\frac{2x}{\tau }z^1+\frac{1}{\tau }z^2\right)\hfill \\ \psi =3_{k0}\left(\frac{3x^2}{\tau }z^1\frac{3x}{\tau }z^2\right)^k\left(\frac{2t_1xt_3}{\tau }z^1\frac{t_1}{\tau }z^2\right).\hfill \end{array}$$ ### Acknowledgments We would like to thank F. Magri for useful comments and suggestions. This work was partially supported by the GNFM branch of the Istituto Nazionale di Alta Matematica, and by the Italian MURST, under the Cofin99 project Geometry of Integrable Systems.
warning/0001/hep-ph0001275.html
ar5iv
text
# Muon-proton Colliders: Leptoquarks, Contact Interactions and Extra Dimensions 1footnote 11footnote 1Invited talk at the 5th International Conference on the Physics Potential and Development of 𝜇⁺⁢𝜇⁻ Colliders, San Francisco CA, December 1999. Work supported by DOE. ## Introduction The R&D rd ; fmc of the muon collider is well underway. The First Muon Collider (FMC) will have a 200 GeV muon beam on a 200 GeV anti-muon beam, which could possibly be at the Fermilab fmc . With the existing Tevatron proton beam the muon-proton collision becomes a possible option. It would be a 200 GeV $``$ 1 TeV $`\mu p`$ collider. The existing lepton-proton collider is the $`ep`$ collider at HERA. Lepton-proton colliders have been proved to be successful by the physics results from HERA. In this work, we shall discuss the physics potential of the $`\mu p`$ colliders at various energies and luminosities. Other $`\mu p`$ colliders that we consider in this study are summarized in Table 1. The nominal yearly luminosity of the 200 GeV $``$ 1 TeV $`\mu p`$ collider is about 13 fb<sup>-1</sup>. Luminosities for other designs are roughly scaled by one quarter power of the muon beam energy and given in Table 1. ## Physics Potential The physics opportunities of $`\mu p`$ colliders are similar to those of $`ep`$ colliders, but the sensitivity reach might be very different, which depends on how precise the particles can be identified and measured in $`ep`$ and $`\mu p`$ environments. Similar to $`ep`$ colliders the proton structure functions can be measured to very large $`Q^2`$ and small $`x`$ in $`\mu p`$ colliders of higher energies. At the 200 GeV $``$ 1 TeV $`\mu p`$ collider the $`Q^2`$ can be measured up to $`10^6`$ GeV<sup>2</sup>. In addition, QCD studies, search for supersymmetry and other exotic particles can also be carried out. Here we concentrate on leptoquarks, leptogluons, $`R`$-parity violating squarks, $`\mu `$-$`q`$ contact interactions, and the large extra dimensions. The goal here is to estimate the sensitivity reach for these new physics at various energies and luminosities. ### Leptoquarks and Leptogluons The second generation leptoquarks made up of a muon and a charm or strange quark are particularly interesting at the $`\mu p`$ collider because they can be directly produced in the $`s`$-channel processes, $`\mu ^\pm c(s)L_{\mu c}(L_{\mu s})`$. It is conventional to assume no inter-generational mixing in order to prevent the dangerous flavor-changing neutral currents. The production cross section of the leptoquark in $`\mu p`$ collisions is $$\sigma =\frac{\pi \lambda ^2}{2s}q(x,Q^2)\times (J+1),$$ (1) where $`\lambda `$ is the coupling constant and $`J`$ is the spin of the leptoquark. On the other hand, a leptogluon has a spin of either $`1/2`$ or $`3/2`$, a lepton quantum number (in this case it is the muon), and a color quantum number (the same as gluon.) The interaction for a spin $`1/2`$ leptogluon is given by $$=g_s\frac{M_{L_{\mu g}}}{2\mathrm{\Lambda }_{\mu g}^2}\overline{L_{\mu g}^a}\sigma ^{\mu \nu }\mu G_{\mu \nu }^b\delta _{ab}+\mathrm{h}.\mathrm{c}.,$$ (2) where $`\mathrm{\Lambda }_{\mu g}`$ is the scale that determines the strength of the interaction. The leptogluon can also be produced in the $`s`$-channel and the production cross section is $$\sigma =\frac{4\pi ^2\alpha _s}{s}\left(\frac{M_{L_{\mu g}}^2}{\mathrm{\Lambda }_{\mu g}^2}\right)^2g(x,Q^2),$$ (3) where $`g(x,Q^2)`$ is the gluon luminosity. The $`R`$-parity violating squarks can be considered special scalar leptoquarks that are the SUSY partners of quarks. The cross section for $`\mu ^+p\stackrel{~}{t}_L`$ is given by $$\sigma _{\stackrel{~}{t}_L}=\frac{\pi |\lambda _{231}^{}|^2}{4s}d\left(\frac{m_{\stackrel{~}{t}_L}^2}{s},Q^2=m_{\stackrel{~}{t}_L}^2\right),$$ (4) where $`d`$ is the down-quark luminosity. The above formula can be easily modified to the production of other squarks with the corresponding subscripts in $`\lambda ^{}`$ and parton functions. If kinematically allowed the leptoquarks, leptogluons, and the $`R`$-parity violating squarks are produced in the $`s`$-channel and thus give rise to a spectacular enhancement in a single bin of the invariant mass $`M`$ distribution or the $`x=M^2/s`$ distribution. ### Contact Interactions The effective four-fermion contact interactions can arise from fermion compositeness or exchanges of heavy particles like heavy $`Z^{}`$, heavy leptoquarks, or other exotic particles. The conventional Lagrangian for $`llqq`$ ($`l=e,\mu `$) contact interactions has the form me $`L_{NC}`$ $`=`$ $`{\displaystyle \underset{q}{}}[\eta _{LL}\left(\overline{l_L}\gamma _\mu l_L\right)\left(\overline{q_L}\gamma ^\mu q_L\right)+\eta _{RR}\left(\overline{l_R}\gamma _\mu l_R\right)\left(\overline{q_R}\gamma ^\mu q_R\right)`$ (5) $`+\eta _{LR}\left(\overline{l_L}\gamma _\mu l_L\right)\left(\overline{q_R}\gamma ^\mu q_R\right)+\eta _{RL}\left(\overline{l_R}\gamma _\mu l_R\right)\left(\overline{q_L}\gamma ^\mu q_L\right)],`$ where $`\eta _{\alpha \beta }^{lq}=ϵ4\pi /\mathrm{\Lambda }_{\alpha \beta }^{lq}{}_{}{}^{2}`$. We introduce the reduced amplitudes $`M_{\alpha \beta }^{\mu q}`$, where the subscripts label the chiralities of the initial lepton ($`\alpha `$) and quark ($`\beta `$). The SM tree-level reduced amplitudes for $`\mu q\mu q`$ are $$M_{\alpha \beta }^{\mu q}(\widehat{t})=\frac{e^2Q_q}{\widehat{t}}+\frac{e^2}{\mathrm{sin}^2\theta _\mathrm{w}\mathrm{cos}^2\theta _\mathrm{w}}\frac{g_\alpha ^\mu g_\beta ^q}{\widehat{t}m_Z^2},\alpha ,\beta =L,R.$$ (6) The new physics contributions to $`M_{\alpha \beta }^{\mu q}`$ from the $`\mu \mu qq`$ contact interactions are $`\mathrm{\Delta }M_{\alpha \beta }^{\mu q}=\eta _{\alpha \beta }^{\mu q}`$. The differential cross sections are given by me $`{\displaystyle \frac{d\sigma (\mu ^+p)}{dxdy}}`$ $`=`$ $`{\displaystyle \frac{sx}{16\pi }}\{u(x,Q^2)\left[\right|M_{LR}^{\mu u}|^2+|M_{RL}^{\mu u}|^2+(1y)^2\left(\right|M_{LL}^{\mu u}|^2+\left|M_{RR}^{\mu u}|^2\right)]`$ (7) $`+d(x,Q^2)\left[\right|M_{LR}^{\mu d}|^2+|M_{RL}^{\mu d}|^2+(1y)^2\left(\right|M_{LL}^{\mu d}|^2+\left|M_{RR}^{\mu d}|^2\right)]\}`$ $`{\displaystyle \frac{d\sigma (\mu ^{}p)}{dxdy}}`$ $`=`$ $`{\displaystyle \frac{sx}{16\pi }}\{u(x,Q^2)\left[\right|M_{LL}^{\mu u}|^2+|M_{RR}^{\mu u}|^2+(1y)^2\left(\right|M_{LR}^{\mu u}|^2+\left|M_{RL}^{\mu u}|^2\right)]`$ (8) $`+d(x,Q^2)\left[\right|M_{LL}^{\mu d}|^2+|M_{RR}^{\mu d}|^2+(1y)^2\left(\right|M_{LR}^{\mu d}|^2+\left|M_{RL}^{\mu d}|^2\right)]\}.`$ ### Model of extra dimensions Arkani-Hamed, Dimopoulos and Dvali add proposed that in the extra dimensions gravity is free to propagate while the SM particles are restricted to a 3-D-brane. The size of the extra dimensions is postulated to be as large as mm to solve the hierarchy problem by bringing the effective Planck scale $`M_S`$ down to TeV. It implies a new gravity interaction for the graviton in the bulk. In our $`3+1`$ dimensional point of view, the graviton behaves as a tower of closely-spaced Kaluza-Klein states. Each state still couples to the SM particles with a normal gravitational strength of order of $`1/M_{\mathrm{Pl}}`$ but, however, there are a huge number of such states. Collectively, the overall coupling strength becomes of order of $`1/M_S`$. In the presence of the new interaction the double differential cross section is given by cheung . $`{\displaystyle \frac{d^2\sigma (\mu ^+p)}{dxdy}}`$ $`=`$ $`{\displaystyle \frac{sx}{16\pi }}{\displaystyle \underset{q}{}}f_q(x)\{(1y)^2(|M_{LL}|^2+|M_{RR}|^2)+|M_{LR}|^2+|M_{RL}|^2`$ (9) $`+{\displaystyle \frac{\pi ^2}{2}}(sx)^2\left({\displaystyle \frac{}{M_S^4}}\right)^2(3264y+42y^210y^3+y^4)`$ $`+2\pi e^2Q_eQ_q\left({\displaystyle \frac{}{M_S^4}}\right){\displaystyle \frac{(2y)^3}{y}}`$ $`+{\displaystyle \frac{2\pi e^2}{\mathrm{sin}^2\theta _\mathrm{w}\mathrm{cos}^2\theta _\mathrm{w}}}\left({\displaystyle \frac{}{M_S^4}}\right)(sx){\displaystyle \frac{1}{Q^2M_Z^2}}[g_a^eg_a^q(6y6y^2+y^3)`$ $`+g_v^eg_v^q(y2)^3]\}`$ $`+`$ $`{\displaystyle \frac{\pi }{2}}f_g(x)(sx)^3\left({\displaystyle \frac{}{M_S^4}}\right)^2(1y)(y^22y+2).`$ Unlike the leptoquarks, the contact interactions and the extra dimensions do not enhance the cross section in a single invariant mass bin, instead, they enhance the cross section at large $`Q^2`$. ## Sensitivity Reach The 95% sensitivity reach on the contact interactions and extra dimensions are calculated as follows. We use the the 2-dimensional $`x`$-$`y`$ distribution to calculate the sensitivity to these new interactions, so as to maximize the sensitivity greg . We divide the $`x`$-$`y`$ plane ($`0.05<x<0.95`$ and $`0.05<y<0.95`$) into a grid. We calculate the number of events predicted by the standard model in each bin with an efficiency of 0.8. We then follow the Monte Carlo approach in Ref. greg . The sensitivity reach on the contact interaction scales $`\mathrm{\Lambda }_{\alpha \beta }^{\mu q}`$ is tabulated in Table 2. The maximum reach of $`\mathrm{\Lambda }`$ at each center-of-mass energy roughly scales as $`\mathrm{\Lambda }40\sqrt{s}`$. The effect of luminosity on $`\mathrm{\Lambda }`$ is rather small: $`\mathrm{\Lambda }`$ only scales as the $`1/4`$th power of the luminosity. The sensitivity reach on the effective Planck scale $`M_S`$ for the model of large extra dimensions is tabulated in Table 3. To estimate the sensitivity reach for $`R`$-parity violating squarks, leptoquarks and leptogluons with a mass $`m`$, we assume the enhancement in cross section is in the mass bin of $`(0.9m,\mathrm{\hspace{0.33em}1.1}m)`$. We calculate the number of events predicted by the standard model in this bin with an efficiency of 0.8, call it $`n^{\mathrm{sm}}`$. Then we use the poisson statistics to estimate the $`n^{\mathrm{th}}`$ that $`n^{\mathrm{sm}}`$ can fluctuate to at the 95% CL. Once the $`n^{\mathrm{th}}`$ is obtained the coupling constant $`\lambda `$ or the leptogluon scale $`\mathrm{\Lambda }_{\mu g}`$ can be determined. These results are tabulated in Tables 4 to 6.
warning/0001/nlin0001054.html
ar5iv
text
# 1 Introduction ## 1 Introduction In recent years there has been a growing interest in discrete analogues of the famous Painlevé equations, i.e. nonlinear nonautonomous ordinary difference equations tending to the continuous Painlevé equations in a well-defined limit and which are integrable in their own right, cf. . Even though the qualitative features of the solutions of these systems are not yet fully understood, nonetheless in most of the known examples the main ingredients of their integrability have been exhibited. Recently, a classification of continuous as well as discrete Painlevé equations in terms of the root systems associated with affine Weyl groups, has been proposed on the basis of the singularities of the rational surfaces of their initial conditions and their blowings-up, cf. . In a recent paper, , we established a connection between the continuous Painlevé VI (P<sub>VI</sub>) equation and a non-autonomous ordinary difference equation depending on four arbitrary parameters. This novel example of a discrete Painlevé equation arises on the one hand as the nonlinear addition formula for the P<sub>VI</sub> transcendents, in fact what is effectively a superposition formula for its Bäcklund-Schlesinger transforms, on the other hand from the similarity reduction on the lattice (cf. ), of a system of partial difference equations associated with the lattice KdV family. In subsequent papers, , some more results on these systems were established, namely the existence of the Miura chain and the discovery of a novel Schwarzian PDE generating the entire (Schwarzian) KdV hierarchy of nonlinear evolution equations and whose similarity reduction is exactly the P<sub>VI</sub> equation, this being the first example of an integrable scalar PDE that reduces to full P<sub>VI</sub> with arbitrary parameters. In the present note we extend these results to multi-dimensional systems associated with higher-order generalisations of the P<sub>VI</sub> equation. Already in we noted that the similarity reduction of the lattice KdV system could be generalised in a natural way to higher-order differential and difference equations, without, however, clarifying in detail the nature of such equations. What we will argue here is that, in fact, such equations constitute what one could call the Painlevé VI hierarchy and its discrete counterpart. Whilst the idea of constructing hierarchies of Painlevé equations by exploiting the similarity reductions of hierarchies of nonlinear evolution equations of KdV type is at least two decades old, cf. , the issue has gained renewed interest in recent years, cf. e.g. , because of the hypothetical possibility that these hierarchies of higher-order Painlevé equations yield new transcendents. Evidence to that effect might be given by the asymptotic analysis of the higher-order equations, since they seem to be governed by hyper-elliptic functions rather than elliptic ones as is the case for the original Painlevé equations, . Most of the existing results on hierarchies of discrete and continuous Painlevé equations are restricted to the examples of P<sub>I</sub> and P<sub>II</sub> hierarchies, since only in these cases it is clear what hierarchies of nonlinear evolution equations should be taken as the starting point for their construction. In the case of the other Painlevé equations, notably P<sub>VI</sub>, it has been less clear what to take as a starting point for the construction of its hierarchy. With the results of we are now well-equipped to tackle this problem, and in the present paper we outline the basic construction of the equations in the discrete as well as continuous P<sub>VI</sub> hierarchy. In fact, we shall demonstrate that the lattice KdV system can be naturally embedded in a multidimensional lattice system achieving the higher-order reductions by including more terms in the relevant similarity constraint which provokes the coupling between the various lattice directions. It should be noted that in a sense higher-order P<sub>VI</sub> systems already were constructed by R. Garnier in his celebrated paper of 1912, , extending the original approach of R. Fuchs who was the first in to find P<sub>VI</sub> arising from the isomonodromic deformation of a second-order linear differential equation. We will conclude our paper with a discussion of these Garnier systems, which in view of the recent interest in algebraic solutions of P<sub>VI</sub>, cf. e.g. -, deserve in our opinion some renewed attention. ## 2 The Discrete PVI Hierarchy In , following earlier work e.g. , cf. also , a coherent framework was developed in which the similarity reduction of both discrete as well as continuous equations associated with the lattice KdV family were treated. Surprisingly, from these reductions the full P<sub>VI</sub> equation for arbitrary parameters emerged together with a four-parameter discrete equation, i.e. a discrete Painlevé equation. From the treatment of it was evident how to to extend the lattice system of partial difference equations and their similarity constraints leading to the reductions to P<sub>VI</sub>. Here we describe explicitely this higher-dimensional lattice system and discuss their explicit reductions. The lattice KdV family of equations contains many related equations such as the lattice Schwarzian KdV, the lattice modified KdV (mKdV) and the actual lattice KdV equations. We concentrate here on one member of this family only, namely the lattice mKdV equation: $$pv_{n,m}v_{n,m+1}+qv_{n,m+1}v_{n+1,m+1}=qv_{n,m}v_{n+1,m}+pv_{n+1,m}v_{n+1,m+1}$$ (2.1) cf. e.g. , with discrete independent variables $`n,m`$ and depending on additional parameters of the equation $`p,q`$, i.e. the lattice parameters. As was pointed out earlier, cf. , the lattice equation (2.1) actually represents a compatible parameter-family of partial difference equations: namely, we can embed the equation (2.1) into a multidimensional lattice by imposing a copy of (2.1) with different parameters on any two-dimensional sublattice, identifying each lattice direction with a corresponding lattice parameter $`p_i`$ in which direction the sites are labelled by discrete variables $`n_i`$ (noting that these are not necessarily integers, but shift by units, i.e. $`n_i\theta _i+`$, $`\theta _i`$). Thus, combining two different lattice directions, labelled by $`(i,j)`$ we can write the lattice equation (2.1) on the corresponding sublattice as $$p_ivv^j+p_jv^jv^{ij}=p_jvv^i+p_iv^iv^{ij}$$ (2.2) in which we use the right superscripts $`i,j`$ to denote the shifts in the corresponding directions, whereas we will use left subscripts $`i,j`$ denote shifts in the reverse direction, i.e. $$v=v(𝒏;𝒑),v^j=T_jv(𝒏;𝒑)=v(𝒏+𝒆_j;𝒑),_jv=T_j^1v(𝒏;𝒑)=v(𝒏𝒆_j;𝒑),$$ where $`𝒏`$ denotes the vector of the discrete variables $`n_i`$, for all lattice directsion labelled by $`i`$, each corresponding to the component $`p_i`$ of the vector $`𝒑`$ of lattice parameters. We use the vector $`𝒆_j`$ to denote the vector with single nonzero entry equal to unity in its $`j^{\mathrm{th}}`$ component. The consistency of the lattice equation (2.2) along the multi-dimensional lattice follows from the diagram of Figure 1: considering the three-dimensional sublattice with elementary directions $`\{𝒆_1,𝒆_2,𝒆_3\}`$ then on each elementary cube in this lattice the iteration of initial data proceeds along the six faces of this cube, on each of which we have an equation of the form (2.2). Thus, starting from initial data $`v`$, $`v^1`$, $`v^2`$, $`v^3`$ we can then uniquely calculate the values of $`v^{12}`$, $`v^{13}`$ and $`v^{23}`$ by using the equation. However, proceeding further there are in principle three different ways to calculate the value of $`v^{123}`$, unless the equation satisfies (as is the case for the equation (2.2)) the special property that these three different ways of calculating this point actually lead to one and the same value. It is indeed at this point that the consistency of the embedding of the lattice MKdV into the multidimensional lattice is tested. In fact, by a straightforward calculation we find that this value is given by $$v^{ijk}=\frac{(p_i^2p_k^2)p_jv^iv^k+(p_j^2p_i^2)p_kv^jv^i+(p_k^2p_j^2)p_iv^jv^k}{(p_i^2p_k^2)p_jv^j+(p_j^2p_i^2)p_kv^k+(p_k^2p_j^2)p_iv^i},i,j,k=1,2,3$$ (which is clearly invariant for any permutation of the labels $`ijk`$), independent of the way in which we calculate this value! Thus, the equation (2.2) can be simultaneously imposed on functions $`v(n_1,n_2,n_3,\mathrm{})`$ of the lattice sites. This is precisely the discrete analogue of the hierarchy of commuting higher-order flows of the (modified) KdV equation! As a consequence of this compatibility we will call the system (2.2) a holonomic system of partial difference equations. Now, we turn to the issue of the symmetry reduction of the multidimensional lattice in the sense of . It follows from the general framework of that the similarity constraint for the multidimensional lattice MKdV system is as follows: $$\underset{i}{}n_ia_i=\mu \nu ,\nu =\lambda (1)^{_in_i},$$ (2.3) $`\mu `$ and $`\lambda `$ being constants, and in which the variables $`a_i`$ are given by $$a_i\frac{v^i_iv}{v^i+_iv}.$$ (2.4) The sum in (2.3) is over all the $`i`$ labelling the lattice directions, the choice of which decides the order of the reduction. To analyse the reduction we need a number of relations for the objects $`a_i`$ which follow from (2.2), namely $`1+a_j^i`$ $`=`$ $`{\displaystyle \frac{(p_iX_{ij}p_j)(a_j+1)+2p_j}{p_ix_{ij}+p_j}},ij`$ (2.5) $`a_i`$ $`=`$ $`{\displaystyle \frac{p_j{}_{i}{}^{}X_{ij}^{}X_{ij}+p_i(_iX_{ij}X_{ij})p_j}{p_j{}_{i}{}^{}X_{ij}^{}X_{ij}p_i(_iX_{ij}+X_{ij})+p_j}}`$ (2.6) $`=`$ $`{\displaystyle \frac{p_j{}_{i}{}^{}x_{ij}^{}x_{ij}p_i(_ix_{ij}x_{ij})+p_j}{p_j{}_{i}{}^{}x_{ij}^{}x_{ij}+p_i(_ix_{ij}+x_{ij})+p_j}},ij`$ in terms of the following variables: $`x_{ij}{\displaystyle \frac{v}{v^{ij}}}`$ , $`{}_{i}{}^{}x_{ij}^{}T_i^1x_{ij}={\displaystyle \frac{{}_{i}{}^{}v}{v^j}},`$ (2.7) $`X_{ij}{\displaystyle \frac{v^i}{v^j}}`$ , $`{}_{i}{}^{}X_{ij}^{}T_i^1X_{ij}={\displaystyle \frac{v}{{}_{i}{}^{}v_{}^{j}}}.`$ (2.8) The variables $`x_{ij}=x_{ji}`$ and $`X_{ij}=1/X_{ji}`$ are not independent, but related via: $$X_{ij}=\frac{p_ix_{ij}+p_j}{p_jx_{ij}+p_i}x_{ij}=\frac{p_iX_{ij}+p_j}{p_jX_{ij}p_i},$$ (2.9) as well as $$\frac{T_i^1x_{ij}}{X_{ij}}=\frac{1a_i}{1+a_i}.$$ (2.10) We note that since the left-hand side of (2.6) depends only on the label $`i`$ but not on $`j`$, for fixed $`i`$ this represents a set of $`N2`$ coupled first-order ordinary difference equations with respect to the shift in the discrete variable $`n_i`$ between the $`N1`$ variables $`X_{ij}`$, $`ji`$. Furthermore, the relations (2.5), for the same fixed label $`i`$, provide us with a set of $`N1`$ first-order relations between the variables $`a_j`$, $`ji`$, and thus together with the similarity constraint (2.3) where $`a_i`$ is substituted by (2.6) we obtain a set of $`2(N1)`$ first-order nonlinear ordinary difference equations for the $`2(N1)`$ variables $`X_{ij}`$, $`a_j`$, $`ji`$, which together form our higher-order discrete system. In the next section we will explicitely disentangle this coupled system in the cases $`N=2`$ and $`N=3`$. The continuous equation for the PVI hierarchy derive from the differential equations with respect to the lattice parameters $`p_i`$, which read: $$p_i\frac{}{p_i}\mathrm{log}v=n_ia_i.$$ (2.11) It can be shown that the differential relations (2.11) are actually compatible not only amongst themselves, but also with the the discrete equations on the lattice (2.2), i.e. the discrete and continuous flows are commuting: $$\frac{}{p_i}\left(\frac{v}{p_j}\right)=\frac{}{p_j}\left(\frac{v}{p_i}\right),\frac{v^i}{p_j}=T_i\left(\frac{v}{p_j}\right).$$ This can actually be demonstrated by explicit calculation exploiting the discrete relations (2.5), but we will not give the details here (which follow closely the pattern of calculations of ). Thus, we have a coherent framework of a large multidimensional system of equations with discrete (in terms of the variables $`n_i`$) as well as continuous (in terms of the parameters $`p_i`$) commuting flows, in terms of which compatible equations of three different types (partial difference, differential-difference and partial differential) figure in one and the same framework: the partial difference equations are precisely the lattice equations (2.2), the differential-difference equations are the relations (2.11), whilst for the partial differential equations in the scheme we refer to our recent paper . Here we will focus now on the reductions under the symmetry constraint (2.3) in order to derive closed-from ODE’s in terms of the lattice parameter $`p_i`$. To make this reduction explicit we use (2.11) in combination with (2.5)-(2.10) to obtain differential relations for the $`a_i`$, namely $$\frac{a_j}{p_i}=\frac{n_ip_j}{p_j^2p_i^2}\left[(1+a_i)(1a_j)X_{ji}(1+a_j)(1a_i)X_{ij}\right],$$ (2.12) as well as the following relations for the reduced variables $`X_{ij}`$ $`\mu +\nu +p_i{\displaystyle \frac{}{p_i}}\mathrm{log}X_{ij}`$ $`=`$ $`n_i𝒳_{ji}a_i+{\displaystyle \underset{ki}{}}n_k𝒳_{ik}a_k`$ $`+n_i{\displaystyle \frac{p_ip_j}{p_i^2p_j^2}}(X_{ji}X_{ij})+{\displaystyle \underset{ki}{}}n_k{\displaystyle \frac{p_kp_i}{p_k^2p_i^2}}(X_{ik}X_{ki})`$ in which we have abbreviated $$𝒳_{ij}\frac{(p_iX_{ij}p_j)(p_jp_iX_{ji})}{p_j^2p_i^2}=𝒳_{ji}.$$ (2.14) Using (2.12) in conjunction with (2) and using the similarity constraint (2.3) to eliminate the $`a_i`$, we obtain a coupled first-order system of differential equations w.r.t. the independent variable $`t_i=p_i^2`$ in terms of the $`2N2`$ variables $`a_k`$, $`X_{ik}`$ , ($`ki`$). Solving the variables $`a_k`$ from the linear system given by the equations (2) and inserting them into (2.12) we obtain a coupled set of second-order nonlinear differential equations for the variables $`X_{ik}`$. ## 3 Special Cases: N=2, N=3 We will now analyse the basic relations of the general framework presented in the previous section in the cases $`N=2,3`$ only in order to arrive at slightly more explicit equations, demonstrating that the reduction leads to ordinary difference equations (in the discrete case) or to ordinary differential equations (in the continuous case). ### N=2: We will be very brief about the two-dimensional case $`N=2`$ which was the main subject of study in the earlier paper . There the compatibility of the similarity constraint and the lattice equation was stated, and the various relations resulting from (2.5) were already written down. Using in this case the slightly simpler notation: $$a_1=a,a_2=b,x_{12}=x,X_{12}=X$$ and using the discrete independent variables $`n_1=n`$, $`n_2=m`$, as well as the lattice parameters $`p_1=p`$, $`p_2=q`$, we derived the second order nonlinear non-autonomous difference equation: $`{\displaystyle \frac{2(n+1)}{1y_{n+1}y_n}}+{\displaystyle \frac{2n}{1y_ny_{n1}}}=\mu +\lambda (1)^n+2n+1+`$ (3.1) $`+`$ $`{\displaystyle \frac{(\mu \lambda (1)^n)(r^21)y_n+r(1y_n^2)\left[(n+\frac{1}{2})(m+\frac{1}{2})(1)^n\right]}{(r+y_n)(1+ry_n)}},`$ (using a slightly different notation from the one of ), where $`r=p/q`$ and where the variables $`y_n`$ are related to the $`X`$ and $`x`$ by the prescription: $`y_{2n}=x(2n)`$ for the even sites, whilst $`y_{2n+1}=1/X(2n+1)`$ for the odd lattice sites (the latter choice being mainly motivated by the wish to cast the equation into a convenient shape). It was pointed out in that whilst a continuum limit of (3.1) yields the P<sub>V</sub> equation, its general solution can be expressed in terms of P<sub>VI</sub> transcendents (noting its dependence on four arbitrary parameters, $`\mu `$, $`\lambda `$, $`r`$ and $`m`$). The continuous equation for the variable $`X`$ in terms of the lattice parameter $`p`$ as independent variable in this case reads: $`p(p^2q^2)^2X(qXp)(pXq){\displaystyle \frac{^2X}{p^2}}=`$ $`={\displaystyle \frac{1}{2}}p(p^2q^2)^2\left[pq(3X^2+1)2(p^2+q^2)X\right]\left({\displaystyle \frac{X}{p}}\right)^2+`$ $`+(q^2p^2)\left[2p^2X(pXq)(qXp)+(q^2p^2)^2X^2\right]{\displaystyle \frac{X}{p}}`$ $`+{\displaystyle \frac{1}{2}}q\left[(\alpha X^2\beta )(pXq)^2(qXp)^2+(p^2q^2)X^2\left((\gamma 1)(qXp)^2(\delta 1)(pXq)^2\right)\right],`$ and it is not difficult to show that this is actually the P<sub>VI</sub> equation through the identification $`w(t)=pX(p)`$, where $`t=p^2`$, and setting $`q=1`$, leading to $`{\displaystyle \frac{d^2w}{dt^2}}=`$ $`{\displaystyle \frac{1}{2}}\left({\displaystyle \frac{1}{w}}+{\displaystyle \frac{1}{w1}}+{\displaystyle \frac{1}{wt}}\right)\left({\displaystyle \frac{dw}{dt}}\right)^2\left({\displaystyle \frac{1}{t}}+{\displaystyle \frac{1}{t1}}+{\displaystyle \frac{1}{wt}}\right){\displaystyle \frac{dw}{dt}}`$ (3.3a) $`+{\displaystyle \frac{w(w1)(wt)}{8t^2(t1)^2}}\left(\alpha \beta {\displaystyle \frac{t}{w^2}}+\gamma {\displaystyle \frac{t1}{(w1)^2}}(\delta 4){\displaystyle \frac{t(t1)}{(wt)^2}}\right),`$ with the identification of the parameters $`\alpha `$,$`\beta `$,$`\gamma `$,$`\delta `$ as follows: $`\alpha =(\mu \nu +mn)^2,\beta =(\mu \nu m+n)^2,`$ (3.3b) $`\gamma =(\mu +\nu mn1)^2,\delta =(\mu +\nu +m+n+1)^2.`$ Eq. (3) is interesting in its own right since it provides us with a covariant way of writing P<sub>VI</sub>, noting its invariance under the transformations: $$nm,pq,X1/X.$$ ### N=3: This first higher-order case deals with the first genuinely multidimensional situation of three two-dimensional sublattices, on each of which a copy of the lattice MKdV equation (2.2) is defined. In addition there is also the similarity constraint (2.3) which couples the three lattice directions. Thus, for the three-dimensional case we have a coupled system of equations whose symbolical representation is shown in figure 2. In the previous section we have already demonstrated the consistency of the three copies of the lattice equation (2.2) amongst themselves. What remains is to investigate the compatability of the lattice equation and the similarity constraint, demonstrating that the determination of the values of the dependent variable $`v`$ by using the lattice equation in all three directions plus the similarity constraint is unique (assuring the single-valuedness of the solution around localised configurations). In Figure 3 we have indicated how the iteration of the system proceeds starting from a given configuration of initial data (located at the vertices indicated by $``$) and moving through the lattice by calculating each point by means of either the lattice equation (points indicated by $``$) or the similarity constraint (points indicated by $`\times `$). The first point where a possible conflict arises, due to the fact that the corresponding values of the dependent variable can be calculated in more than one way, is indicated by $``$. It is at such points that the consistency of the similarity reduction needs to be verified by explicit computation. This has been carried out for this three-dimensional case using MAPLE. Obviously, the iteration involves too many steps and the expressions soon become too large to reproduce here. In order to analyse the explicit reduction in this case, we redefine the following objects $`a_1=a,a_2=b,a_3=c`$ $`X_{12}=X,X_{13}=Y,𝒳_{12}=𝒳,𝒳_{13}=𝒴`$ using also $`n_1=n`$, $`n_2=m`$, $`n_3=h`$, as well as $`p_1=p`$, $`p_2=q`$ and $`p_3=r`$ to simplify the notation. To start with the continuous equations, fixing the independent variable to be $`p`$ we obtain the following linear system for the quantities $`b`$ and $`c`$ from eq. (2) $`\left[\begin{array}{cc}2𝒳& 𝒳+𝒴\\ 𝒳+𝒴& 2𝒴\end{array}\right]\left[\begin{array}{c}mb\\ hc\end{array}\right]=\left[\begin{array}{c}\mu +\nu +p\frac{}{p}\mathrm{log}X\\ \mu +\nu +p\frac{}{p}\mathrm{log}Y\end{array}\right]+`$ (3.10) $`\left[\begin{array}{cc}(\mu \nu )𝒳+(n+m)\frac{pq}{p^2q^2}(\frac{1}{X}X)+h\frac{pr}{p^2r^2}(\frac{1}{Y}Y)& \\ (\mu \nu )𝒴+(n+h)\frac{pr}{p^2r^2}(\frac{1}{Y}Y)+m\frac{pq}{p^2q^2}(\frac{1}{X}X)& \end{array}\right]`$ (3.13) where we have used the similarity constraint to elininate the quantity $`a`$. Furthermore, from (2.12) we obtain the differential relations $`{\displaystyle \frac{(mb)}{p}}`$ $`=`$ $`{\displaystyle \frac{mq}{q^2p^2}}[(n+\mu \nu mbhc)(1b){\displaystyle \frac{1}{X}}`$ (3.14a) $`(1+b)(n\mu +\nu +mb+hc)X]`$ $`{\displaystyle \frac{(hc)}{p}}`$ $`=`$ $`{\displaystyle \frac{hr}{r^2p^2}}[(n+\mu \nu mbhc)(1c){\displaystyle \frac{1}{Y}}`$ (3.14b) $`(1+c)(n\mu +\nu +mb+hc)Y].`$ Solving $`b`$ and $`c`$ from the linear system (3.13), and substituting the results in the differential relations (3.14a) and (3.14b), we obtain two coupled second-order nonlinear ODE’s depending effectively on six free parameters, namely $`\mu `$, $`\nu `$, $`n`$, $`m`$, $`h`$ and $`q/r`$. Alternatively, we can derive a system of second-order ordinary difference equations by fixing one of the discrete variables, say $`n=n_1`$, and using the relations (2.5) to obtain the equations $`(pXq)b+pX+q`$ $`=`$ $`{\displaystyle \frac{(q^2p^2)X}{qXp}}(\stackrel{~}{b}+1)`$ (3.15a) $`(pYr)c+pY+r`$ $`=`$ $`{\displaystyle \frac{(r^2p^2)Y}{rYp}}(\stackrel{~}{c}+1),`$ (3.15b) where the tilde denotes the shift in the lattice direction associated with the variable $`n`$. Using the similarity constraint $$na+mb+hc=\mu \nu ,\nu =\lambda (1)^{n+m+h}$$ (3.16) to eliminate the variables $`c`$, we obtain the following linear system in terms of $`\stackrel{~}{b}`$ and $`b`$ $`\left[\begin{array}{cc}(q^2p^2)X& (pXq)(qXp)\\ m(r^2p^2)Y& m(pYr)(rYp)\end{array}\right]\left[\begin{array}{c}\stackrel{~}{b}\\ b\end{array}\right]=`$ (3.21) $`\left[\begin{array}{c}(pX+q)(qXp)(q^2p^2)X\\ (rYp)((pYr)(\mu \nu na)+h(pY+r))(r^2p^2)Y(h+\mu +\nu (n+1)\stackrel{~}{a})\end{array}\right]`$ (3.24) where the $`a`$ and $`\stackrel{~}{a}`$ can be expressed in terms of $`X`$ and $`Y`$ by $$a=\frac{qX\begin{array}{c}X\hfill \\ \mathrm{~}\hfill \end{array}+p(\begin{array}{c}X\hfill \\ \mathrm{~}\hfill \end{array}X)q}{qX\begin{array}{c}X\hfill \\ \mathrm{~}\hfill \end{array}p(\begin{array}{c}X\hfill \\ \mathrm{~}\hfill \end{array}+X)+q}=\frac{rY\begin{array}{c}Y\hfill \\ \mathrm{~}\hfill \end{array}+p(\begin{array}{c}Y\hfill \\ \mathrm{~}\hfill \end{array}Y)r}{rY\begin{array}{c}Y\hfill \\ \mathrm{~}\hfill \end{array}p(\begin{array}{c}Y\hfill \\ \mathrm{~}\hfill \end{array}+Y)+q},$$ (3.25) (where the undertilde denotes the backward shift with respect to the discrete variable). The system of equations (3.15), (3.16) and (3.25) – or, equivalently, (3.21) together with (3.25) leads in principle to a fourth order ordinary difference equation in one variable. In fact, solving $`b`$ and $`\stackrel{~}{b}`$ from (3.21) and then eliminating $`b`$ altogether by a shift in the independent variable $`n`$ we get a coupled system containing one equation in terms of $`X`$, $`\stackrel{~}{X}`$, $`\stackrel{~}{\stackrel{~}{X}}`$, $`\begin{array}{c}X\hfill \\ \mathrm{~}\hfill \end{array}`$ and $`Y`$, and the equation (3.25) which is first order in the both $`X`$ and $`Y`$ with respect to the shift in the variable $`n`$. This system of equations depends effectively on six free parameters, namely $`\mu `$, $`\nu `$, $`m`$, $`h`$, $`q/p`$ and $`r/p`$. ## 4 Isomonodromic Deformation Problem The isomonodromic deformation problem for the multidimensional lattice system is of Schlesinger type, . In the two-dimensional case it was already presented in for special values of the parameters $`\mu `$, $`\nu `$, cf. also for the general parameter case. The extension from the two-dimensional to the multidimensional lattice is immediate: one only needs to introduce additional terms of similar form for each additional lattice direction. Thus, the Lax representation consists on the one hand of the linear shifts on the lattice of the form $$\psi ^i(\kappa )=T_i\psi (\kappa )=L_i(\kappa )\psi (\kappa ),$$ (4.1) in which $`\kappa `$ is a spectral parameter, and where the Lax matrices $`L_i`$ are given by $$L_i(\kappa )=\left(\begin{array}{cc}p_i& v^i\\ \frac{\kappa }{v}& p_i\frac{v^i}{v}\end{array}\right),$$ (4.2) leading to the Lax equations $$L_i^jL_j=L_j^iL_i$$ (4.3) which lead to a copy of the lattice MKdV equation on each two-dimensional sublattice labelled by the indices $`(i,j)`$. On the other hand we have the linear differential equation for $`\psi (\kappa )`$ with respect to its dependence on the spectral variable $`\kappa `$ $`\kappa {\displaystyle \frac{d}{d\kappa }}\psi (\kappa )`$ $`=`$ $`{\displaystyle \frac{1}{2}}\left(\begin{array}{cc}(1+\mu )& 0\\ 0& \lambda (1)^{_in_i}+_in_i\end{array}\right)\psi (\kappa )`$ (4.9) $`+{\displaystyle \underset{i}{}}{\displaystyle \frac{n_iv}{v^i+_iv}}\left(\begin{array}{cc}0& v^i\\ 0& p_i\end{array}\right)T_i^1\psi (\kappa ),`$ the compatibility of which with (4.1) leads to the similarity constraint (2.3). In addition, we have differential equations for $`\psi `$ in terms of its dependence on the lattice parameters $`p_i`$ which are of the form $$\frac{\psi }{p_i}=\frac{n_i}{p_i}\left(\begin{array}{cc}1& 0\\ 0& 0\end{array}\right)\psi +\frac{2n_iv}{v^i+_iv}\left(\begin{array}{cc}0& \frac{1}{p_i}v^i\\ 0& 1\end{array}\right)T_i^1\psi ,$$ (4.10) for each of the variable $`p_i`$. It is the variables $`t_i=p_i^2`$ that play the role as independent variables in the continuous PVI hierarchy. The elimination of the back-shifted vectors $`T_i^1\psi `$ by using the inverse of the Lax relations (4.1) lead to the following linear differential equation for $`\psi `$ $$\frac{\psi }{\kappa }=\left(\frac{A_0}{\kappa }+\underset{i}{}\frac{A_i}{\kappa t_i}\right)\psi $$ (4.11) thus leading to the problem in the Schlesinger form, with regular singularities at $`0,\mathrm{},\{t_i\}`$. The matrices $`A_0`$ and $`A_i`$ are given by $`A_0={\displaystyle \frac{1}{2}}\left(\begin{array}{cc}(\mu +1)& _i\frac{n_i}{p_i}(1a_i)v^i\\ 0& \lambda ^{_in_i}+_in_ia_i\end{array}\right),`$ $`A_i=n_i\left(\begin{array}{cc}\frac{1}{2}(1+a_i)& \frac{1}{2p_i}v^i(1a_i)\\ \frac{p_i}{2v^i}(1+a_i)& \frac{1}{2}(1a_i)\end{array}\right).`$ The continuous isomonodromic deformation is provided by the linear differential equations in terms of the lattice parameters, namely $$\frac{\psi }{t_i}=\left(P_i\frac{A_i}{\kappa t_i}\right)\psi $$ (4.14) where $$P_i=\frac{n_i}{2p_i}\left(\begin{array}{cc}\frac{1}{p_i}a_i& 0\\ \frac{1}{v^i}(1+a_i)& 0\end{array}\right).$$ Eq. (4.14) is not quite in standard form, and we need to apply a gauge transformation of the form $$\overline{\psi }V\psi ,V=\left(\begin{array}{cc}1/v& 0\\ U/v& 1\end{array}\right),$$ (4.15) to remove the term with $`P_i`$, where the auxiliary variable $`U`$ obeys an interesting set of equations by itself (in fact this is the object obeying the lattice KdV system of equations) , cf. , but we will not give any details here. With this gauge, the continuous isomonodromic deformation (4.14) adopts the standard form $$\frac{\overline{\psi }}{t_i}=\frac{\overline{A}_i}{\kappa t_i}\overline{\psi },\overline{A}_i=VA_iV^1,$$ (4.16) whilst the discrete isomonodromic is readily obtained from the Lax representation of (4.1). ## 5 Connection with Garnier Systems Interestingly, already M.R. Garnier in his seminal paper of 1912, , embarked on the question of finding higher-order analogues of the PVI equation, adopting the method that was proposed somewhat earlier by R. Fuchs, in , which can be identified with the isomonodromic deformation approach, cf. also . Garnier gave a general construction of such higher-order equations constituting coupled systems of partial differential equations, which are the isomonodromic Garnier systems. As a particular example, he wrote down explicitely in the first higher-order PVI equation in terms of the following coupled system, consisting of the second order ODE in terms of two dependent variables $`w=w(t,s)`$ and $`z=z(t,s)`$ $`{\displaystyle \frac{^2w}{t^2}}`$ $`=`$ $`{\displaystyle \frac{1}{2}}\left({\displaystyle \frac{1}{w}}+{\displaystyle \frac{1}{w1}}+{\displaystyle \frac{1}{wt}}+{\displaystyle \frac{1}{ws}}{\displaystyle \frac{1}{wz}}\right)\left({\displaystyle \frac{w}{t}}\right)^2`$ $`\left({\displaystyle \frac{1}{t}}+{\displaystyle \frac{1}{t1}}+{\displaystyle \frac{1}{ts}}{\displaystyle \frac{1}{tw}}{\displaystyle \frac{1}{tz}}\right){\displaystyle \frac{w}{t}}`$ $`+{\displaystyle \frac{1}{2}}{\displaystyle \frac{w(w1)(ws)(zt)}{z(z1)(zs)(wt)(zw)}}\left({\displaystyle \frac{z}{t}}\right)^2{\displaystyle \frac{wt}{(zt)(zw)}}\left({\displaystyle \frac{w}{t}}\right)\left({\displaystyle \frac{z}{t}}\right)`$ $`+{\displaystyle \frac{2w(w1)(wt)(ws)(zt)^2}{t^2(t1)^2(ts)^2(wz)}}\times `$ $`\times [\alpha +\beta +\gamma +\delta +\kappa +{\displaystyle \frac{7}{4}}{\displaystyle \frac{ts}{z}}{\displaystyle \frac{\alpha +\frac{1}{4}}{w^2}}+{\displaystyle \frac{(t1)(s1)}{(z1)}}{\displaystyle \frac{\beta +\frac{1}{4}}{(w1)^2}}`$ $`+{\displaystyle \frac{t(t1)(ts)}{(zt)}}{\displaystyle \frac{\gamma }{(wt)^2}}+{\displaystyle \frac{s(s1)(st)}{(zs)}}{\displaystyle \frac{\delta }{(ws)^2}}]`$ together with coupled first order PDE’s $`{\displaystyle \frac{t(t1)}{tz}}{\displaystyle \frac{w}{t}}+{\displaystyle \frac{s(s1)}{sz}}{\displaystyle \frac{w}{s}}={\displaystyle \frac{w(w1)}{wz}},`$ (5.1b) $`{\displaystyle \frac{t(t1)}{tw}}{\displaystyle \frac{z}{t}}+{\displaystyle \frac{s(s1)}{sw}}{\displaystyle \frac{z}{s}}={\displaystyle \frac{z(z1)}{zw}}.`$ (5.1c) It should be pointed out that the system consisting of (5.1), (5.1b) and (5.1c) amounts actually to a fourth order ODE in terms of $`w=w(t)`$ only, and as such can be rightly considered to be the first higher-order member of the Painlevé VI hierarchy. In fact, Garnier gave in his paper a number of important assertions: i) that his system of equations is completely integrable<sup>1</sup><sup>1</sup>1Obviously, Garnier’s use of the term integrability was meant here in the precise sense of that of a compatible system, very much in the same sense as the compatibility of the continuous and discrete systems that we have encountered in sections 2 and 3., and that it degenerates (under the autonomous limit) to a system living on the Jacobian of a hypereliptic curve, ii) that the symmetric combinations of the dependent variables of the system, as functions of each one of the essential singularities (i.e. singling out one of the independent variables) are meromorphic in terms this variable except for the fixed critical points which are at 0,1, $`\mathrm{}`$, or at the location of the values of the other independent variables<sup>2</sup><sup>2</sup>2This assertion amounts to the well-known Painlevé property., iii) that for the parameters of the system in general position the symmetric functions of the dependent variables are essentially transcendental functions of the constants of integration (i.e. of the initial data). Subsequent work on the Garnier systems was done mostly by K. Okamoto and his school, cf. e.g. . However, it seems that in most of these works these systems were treated rather as an underdetermined system of PDE’s rather than (as Garnier himself clearly had in mind) as a consist system of ODE’s. Although it is not easy to find the explicit transformation of the lattice system exposed in sections 2 and 3 to the systems that Garnier wrote down, in particular to find the explicit relation between the above system (5.1) and the system consisting of (3.13) and (3.14), it is to be expected that such a mapping exist. The identification is probably easiest to obtain via the transformation of the corresponding Schlesinger type of system as given in section 4 and the linear system that Garnier exploited in . However, the search for such an identification will be left to a future study. Let us finish with some remarks on the relevance of these results for work that is been done in recent years. One of the most exciting developments is the way in which the issue of algebraic solutions of P<sub>VI</sub> have arisen in recent years, e.g. in connection with WDVV equations, Frobenius manifolds and quantum cohomology, cf. e.g. the review . Such algebraic solutions were already known to Picard, Painlevé and Chazy. In fact, in his early paper R. Fuchs obtained a realisation of P<sub>VI</sub> in terms of an elliptic integral, and this realisation was subsequently used by Painlevé in , to derive an elliptic form for the P<sub>VI</sub> equation, a form of the equation that was recently recovered by Manin in <sup>3</sup><sup>3</sup>3We are grateful to R. Conte for pointing out the reference .. The assertions of Garnier in on his generalisation of the Fuchs’ approach might form a starting point for extending this elliptic connection to the Garnier systems, in which case we would expect to be able to find a realisation of those systems in terms of hyperelliptic integrals rather than elliptic ones. This might eventually lead to the construction of algebraic solutions of those systems, possibly in the spirit of the recent papers . It would be of interest to further investigate the role of the discrete systems in connection with the Garnier systems: we expect them to constitute the superposition formulae for the underlying higher root systems of the corresponding affine Weyl groups. Thus, eventually, a geometric interpretation of the Garnier systems and their discrete analogues in the sense of the blowings-up of the corresponding rational surfaces of their initial conditions, along the lines of the recent paper , might be anticipated.
warning/0001/hep-ph0001012.html
ar5iv
text
# Electron - Proton Scattering as a Probe of Nucleon Structure ## I Introduction Recent experiments for electron-proton (ep) scattering, SLAC-E-136 , DESY-HERA-HE , SLAC-E-133 and others , played important role in probing the nucleon structure and to reveal the dynamic mechanism of the electron interactions inside the nucleon bag. Many trials have been done before by a wide variety of empirical and theoretical models, ranging from form factor scaling , and vector meson dominance to quark- parton models and perturbative QCD . An eikonal optical picture \[9-10\] was used, based on multiple scattering of the incident electron with the constituent valance quarks of the target hadron, assuming different forms of the (electron-quark) binary wave functions. In this article we shall deal with the problem of elastic and inelastic collisions of the electron with proton from a different point of view. In the following, we present two pictures for the (ep) scattering. The first is a scattering in view of multi-photon exchange mechanism (MPEM) which is very convenient for the elastic scattering in a wide range of momentum transfer square $`Q^2`$. On the other hand, the multi-peripheral model (MPM) is relevant to the inelastic and deep inelastic scattering and so works properly for the problems of multi particle production. The article structure goes as follows. In section 2, we present the postulates of the MPEM, the formulation and the scenario of the MPM are given in section 3. The Monte Carlo generators are summarized in section 4. Finally results and discussion are given in section 5, followed by conclusive remarks. ## II The multi photon exchange model (MPEM) It is assumed that elastic scattering of electrons on proton proceeds via multi step processes represented by ladder diagrams . We proceed on the bases of the Feynman formalisms, assuming electromagnetic interaction acting at each vertex of the ladder diagram Fig.(1). Expanding the transition matrix $`T`$ of the (ep) scattering in terms of transition ladder diagrams $`T^{(n)}`$ so that, $$T=\underset{i}{\overset{n1}{}}T_i$$ (1) $`\{c^n\}`$ and $`\{T^{(n)}\}`$ are the coefficients of expansion and transition matrices of the adder diagram of order $`n`$. $`T^{(n)}`$ has the form, $$T^{(n)}=\mathrm{}\underset{1}{\overset{n1}{}}dk_j\frac{V_{j+1,j}V_{j,j1}}{kk_j+iϵ}.$$ (2) Where the factor $`\frac{1}{kk_j+iϵ}`$ stands for the Green’s propagator of the virtual intermediate state number $`j`$. $`V_{j,j1}`$ is the transition probability from the state $`j1`$ to $`j`$, so that, $`V_{j,j1}=<\varphi _j|V|\varphi _{j1}>.`$ The screening Coulomb field working at each vertex has the form, $`V=U_0\frac{e^{\alpha r}}{r}`$ so that, each internal integration, $`V_{kk^{}}`$ has the form; $$<k|V|k^{}>=U_0[2\pi ^2(\alpha ^2+|kk^{}|^2)].$$ (3) Hence, a one step ladder diagram has a transition matrix $`T^{(1)}`$, $$T^{(1)}=U_0[2\pi ^2(\alpha ^2+|k_ik_f^{}|^2)]$$ (4) $`k_i`$ and $`k_f`$ are the momentum of the initial and final states. Similarly, the two step and the three step ladders will have the forms, $$T^{(2)}=2\pi ^2U_0^2\frac{d\underset{1}{\overset{\mathrm{\_}}{k}}}{(k_1k^2iϵ)(\alpha ^2+|k_ik_1^{}|^2)(\alpha ^2+|k_1k_f^{}|^2)}$$ (5) $`T^{(3)}={\displaystyle }{\displaystyle }{\displaystyle \frac{(U_0^3/4\pi ^4)d\underset{1}{\overset{\mathrm{\_}}{k}}d\underset{2}{\overset{\mathrm{\_}}{k}}}{(kk_2^2iϵ)(\alpha ^2+|k_2k_1^{}|^2)(\alpha ^2+|k_2k_f^{}|^2)(kk_1^2iϵ)}}`$ $$\frac{1}{(kk_1^2iϵ)(\alpha ^2+|k_ik_1^{}|^2)}$$ (6) The integrals in Eqs.(5,6) are carried out by the Dalitz integrals . The expansion coefficients $`\{c_i\}`$ in Eq.(1) are determined by fitting with the experimental data to estimate the statistical weight factor of each ladder diagram contributing the reaction. According to the usual formalism for the dynamics of the particle reactions , the phase space integral is an integration of the square of the transition matrix $`T(\{k_i\})`$ over a set of allowed values $`\{k_i\}`$. $$I_n(s)=\underset{i}{\overset{n}{}}\frac{d^3k_i}{2E_i}\delta ^4(k_a+k_b\underset{i=1}{\overset{n}{}}k_i)|T\{k_i\}|^2$$ (7) $`k_a,k_b`$ are the momenta of the interacting particles in the initial state, while $`\{k_i\}`$being the momenta of the particles in the final state. For the problem under consideration of the elastic scattering, where only two particles in the final state $`n=2`$, with the normalization convention, the reaction cross section for $`k_a+k_bk_1+k_2`$ is $$\sigma (s)=\frac{1}{8\pi ^2\lambda ^{1/2}(s,m_a^2,m_b^2)}\frac{d^3k_1}{2E_1}\frac{d^3k_2}{2E_2}\delta ^4(k_a+k_bk_1+k_2)|T\{k_1,k_2\}|^2$$ (8) $`\lambda `$ is a standard function defined as,$`\lambda (x,y,z)=(xyz)^24yz`$ . One may immediately write the differential cross section in the center of mass system CMS $$\frac{d\sigma }{d\mathrm{\Omega }}=\frac{1}{64\pi ^2s}\frac{k_1}{k_a}|T|^2$$ (9) It is more convenient to use the invariant cross section, $$\frac{d\sigma }{dt}=\frac{1}{16\pi \lambda (s,m_a^2,m_b^2)}|T|^2,t=Q^2$$ (10) Eqs. (9),(10) as well as Eq.(1) are in a relevant form to compare with the experimental data. ## III The multi peripheral model (MPM) In this section, we shall deal with the problem of the particle production in the (ep) inelastic scattering in view of the multi peripheral collision . A factorizable transition matrix $`T`$ is assumed in the form, $$T=\underset{i}{\overset{n1}{}}T_i$$ (11) Each particle in the final state is produced at a specific peripheral surface as shown by the Feynman diagram Fig.(2) with electromagnetic transition matrix $`T_i`$, may be written in a suitable parametric form, $$T=\frac{1}{\alpha _i+t_i}$$ (12) where t<sub>i</sub> is the four vector momentum transfer square at the $`i^{th}`$ peripheral surface. $`\alpha _i`$ is the electromagnetic peripheral parameter characterizing the surface number $`i`$, and determine to conserve the total energy. The advantage of this technique is to reduce the many body problem into $`(n1)`$ iterative diagrams, each of them has only two particles in the final state. For example, the $`i^{th}`$ diagram has two particles in the final state, the first one is the particle number $`(i+1)`$, and the other one has an effective mass $`M_i`$, equivalent to the rest of the i-particles of the system. The square of the 4-vector momentum transfer $`t_i`$ is kinematically calculated as, $`t_i=(p_a^{(i)}p_1\mathrm{}.p_i)`$ $$t_i=m_{i+1}^2+M_i^22E_a^{(i)}k_i^02P_a^{(i)}K_i\mathrm{cos}\theta _i$$ (13) $`m_{i+1}`$ is the rest mass of the particle number $`(i+1)`$ produced in the $`i^{th}`$ iteration. $`K_i`$ and $`k_i^0`$ are the 3-vector momentum and the total energy of the effective mass $`M_i`$. and are the corresponding figures for the leading particle acting at the $`i^{th}`$ peripheral surface. The recursion relation of $`P_a^{(i)}`$is given by, $$P_a^{(i)}=\lambda ^{1/2}(M_i^2,t_i,m_a^2)/2M_i$$ (14) The leading particle for the first peripheral surface is given by $$P_a=P_a^{(n)}=\lambda ^{1/2}(s,m_a^2,m_b^2)/2\sqrt{s}$$ (15) The multi peripheral parameters $`\{\alpha _i\}`$ play important role in converging the particles in phase space and consequently, control the energy of the particles in final state. So that the values of $`\{\alpha _i\}`$ are adjusted to conserve the total energy. The energy $`E_i`$ of the particle number $`i`$ is related to its rapidity $`y_i`$ through the relation, $`E_i=m_t\mathrm{cosh}(y_i)`$ $$m_t=\sqrt{P_t^2+m_i^2}$$ (16) so that the total energy of the particles in the final state is $$\zeta _i^n=\frac{1}{\sigma _n}m_t\mathrm{cosh}(y)(d\sigma /dy)𝑑y$$ (17) $`\zeta _i^n`$ which is a function of the parameters $`\{\alpha _i\}`$ should be compared with the total center of mass energy$`\sqrt{s}`$ of the initial state. We first start with $`n=2`$ to get $`\alpha _1`$, which is inserted again in the case $`n=3`$ to get $`\alpha _2`$ and so on. These are repeated sequentially to get the values of the rest parameters up to $`\alpha _{n1}`$. The values $`\{\alpha _i\}`$ depend on the particle multiplicity n rather than the energy $`\sqrt{s}`$. The phase space integral $`I_n(s)`$ is then calculated as in Eq.(7) after transforming the integral variables from $`k_i`$ to $`t_i`$, $`I_n(s)={\displaystyle \frac{(2\pi )^{n1}}{2M_n}}{\displaystyle \underset{i=3}{\overset{n}{}}}{\displaystyle \frac{1}{4P_a^{(i)}}}{\displaystyle _{\mu _i}^{M_im_i}}dM_{i1}{\displaystyle _{t_{i1}^{}}^{t_{i1}^+}}{\displaystyle \frac{1}{t_{i1}+\alpha _{i1})^2}}dt_{i1}`$ $$\frac{1}{4P_a^{(2)}}_{t_1^{}}^{t_1^+}\frac{1}{t_1+\alpha _1)^2}𝑑t_1$$ (18) $$\mu _i=\underset{j=1}{\overset{i}{}}m_j$$ (19) $`t_i^\pm `$ are the upper and lower limits of $`t_i`$ corresponding to $`cos(\theta _i)=\pm 1.`$ $$t__i^\pm =m_a^2+m_b^22E_aE_b\pm 2P_aP_b$$ (20) Multiple integrals in Eq.(18) are carried out using a Monte Carlo program, through which all possible distributions of the physical quantities are easily found. ## IV Monte Carlo A Monte Carlo program GENE2 is developed to simulate events according to the multi peripheral diagrams. It includes 3-generators. The generator $`Gn(s)`$ for the multiplicity of particles in the final state, the generator $`GM(s,n)`$ for the invariant masses produced at the $`n1`$ possible peripheral surfaces and finally the dynamic generator $`Gt(s,n,M)`$ which generates the values of the momentum transfer square $`t_i`$ of the $`i^{th}`$ surface according to electromagnetic transition matrix. The program GENE2 executes the 3- generators in ordered sequences, and simulates all the kinematical variables of the n- particles in the final state. A general form of the algorithm used to generate an event x with probability $`f(x)`$ is, $$r=_a^xf(x)𝑑x/_a^bf(x)𝑑x$$ (21) Where $`a`$ and $`b`$ are the boundaries of the physical region at which x is defined. $`r`$ is the default random number with uniform distribution between the limits $`\{0,1\}`$. ### A The multiplicity generator Gn(s) It is assumed that the multiplicity distribution $`P(n)`$ has a Gaussian form , $$P(n)=\frac{1}{\sqrt{2\pi }\sigma }\mathrm{exp}[(n\stackrel{}{n})^2/\sigma ^2]$$ (22) where $`\stackrel{}{n}`$ is the average multiplicity which depends on the center of mass energy $`\sqrt{\sigma }`$ and $`\sigma `$ is the dispersion. $$\stackrel{}{n}=a\mathrm{log}(s)+b,\sigma =c\stackrel{}{n}+d$$ (23) the integration limits extend from zero to infinity. This makes Eq.(21) read; $$r=Erf(n\stackrel{}{n})$$ (24) $`Erf(x)`$ is the error function. The solution $`n`$ of Eq. (24) defines the multiplicity generator of the reaction. ### B The invariant mass generator GM(s,n) According to the multi peripheral diagrams Fig.(2), it is assumed that the peripheral surface number i, produces the particle number $`i+1`$ and the invariant mass $`M_i,`$ equivalent to the effective mass of the system of the rest of the $`i`$particles. The value of $`M_i`$ satisfies the relation, $$m_iM_iM_{i+1}$$ (25) so that the algorithm of the generator GM(s,n) is, $$M_i=\underset{j}{\overset{i}{}}m_j+r(M_{i+1}\underset{j=1}{\overset{i}{}}m_j)$$ (26) For a system of n-particle final state, we start with $`M_n=\sqrt{s}`$, then generate the value of $`M_{n1}`$. Eq.(26) is the recursion relation to generate all values of $`M_i,i=n1,\mathrm{},2`$. ### C The dynamic generator Gt(s,n,M) The algorithm used to generate the values of the square of the momentum transfer $`t_i`$ should simulate a probability distribution which is proportional to the square of the transition matrix $`T_i`$ defined by Eq.(12). Inserting this in Eq.(21), we get the $`t_i`$ generator as, $$t_i=\{r[(t_i^++\alpha _i)^1(t_i^{}+\alpha _i)^1+(t_i^++\alpha _i)^1\}^1\alpha _i$$ (27) ## V Results and discussion ### A The elastic scattering The problem of elastic scattering is treated in this article using the ladder diagrams of multi steps. The transition matrix $`T^{(n)}`$ is calculated for ladder diagrams for $`n=1,2`$ and $`3`$ using the Dalitz and Feynman integrals. The results are demonstrated in Fig.(3). The potential field acting at each vertex of the diagram is assumed to be of electromagnetic nature with screening factor due to the pion current inside the proton target. The screening effect limits the infinite range of the coulomb potential. It is found that the terms of the transition matrix $`T^{(n)}`$ form a converging series. The first term of which decreases rapidly with the momentum transfer square $`t=Q^2`$. while the higher order terms are slowly varying functions. The asymptotic behavior of $`T^{(n)}`$ follows a power law at extremely high energy. The expansion coefficients $`\{c_i\}`$ are determined by the fitting method with the SLAC experimental data in the range $`t13(GeV/c)^2`$. The results are shown in Fig.(4). It is found that only two terms of the series are sufficient to represent the reaction. The maximum probable diagram corresponds to $`n=1`$. On the other hand, the SLAC-E-136 experiment, Fig.(5) in the momentum transfer range $`t331(GeV/c)^2`$ needs more terms of the series, with a most probable diagram corresponds also to $`n=1`$. The values of the expansion coefficients are given in Table (1). Table(1) The branching ratios of multi photon exchanged in ep collisions. Reaction 1-Step 2\_Steps 3-Steps Ref. 91 9 - SLAC-E-136 86 12 2 Figure (4) and (5) show good agreement between the experimental data and the prediction of the multi photon exchange mechanism. It is clear that the reactions of high momentum transfer needs more terms of ladder diagrams. ### B The inelastic scattering. Here we use the iterative multi peripheral diagram as in Fig.(2-a,b). The advantage of this is to simulate a complete inclusive reaction. We used the Monte Carlo program GENE2 to generate the n-particles in the final state. The multi peripheral parameters $`\alpha _i`$, $`i=1,2,\mathrm{},n1`$ are determined to conserve the total center of mass energy $`\sqrt{s}`$. The values of the parameters $`\{\alpha _i\}`$ depend not only on $`\sqrt{s}`$ but also on the degree of peripherality $`i`$. The first few peripheral surfaces possess parameter values which confine the particle production in very narrow cone. The production cone angle gets wider for higher order surfaces. The values of $`\alpha _i`$ are displayed in Fig.(6) as a function of the multiplicity $`n`$, of the number of particle in the final state for electron lab energies $`50,100`$ and $`200GeV`$. In all cases the value of $`\alpha _i`$ decreases slowly with n, up to a critical multiplicity nc after which a sudden drop is observed. A parametric relation is obtained for $`\alpha _i`$ as a function of the electron lab energy $`E`$ and $`n`$ as, $`\alpha _{n1}=(0.2260.015E+0.0003E^2)n+0.441\mathrm{log}(E)2.29,n<n_c.`$ $`\alpha _{n1}=(0.0019E0.8685)\mathrm{exp}[(.000688E+0.327276)n],n>n_c`$ $$n_c=0.4846E+1.0769$$ (28) The critical value $`n_c`$ corresponds to the peripheral surface at which enough energy is transferred, that is sufficient to make phase transition from the nuclear matter to the quark gluon state. The model is applied to the data of the experiment SLAC-E-133 corresponding to electron lab energies $`9.744,12.505,15.730,18.476`$ and $`20.999GeV.`$ Fig.(7) shows the rapidity distributions for particles produced at $`9.744GeV`$ electron lab energy as calculated by (MPM) at peripheral surfaces $`n=2,3,5`$ and $`10`$. The result for $`n=2`$ (two particle final state) shows two non symmetric clear peaks corresponding to forward and backward emission. The electron peak is relatively narrower than the proton one . This case represents the most peripheral collision with low momentum transfer. As the multiplicity increases the two peaks get broader and interfere through each other. They completely interfere and form only one peak at the case of high multiplicity $`n=10`$ corresponding to high momentum transfer. This represents the most central collision. i.e. collision between the electron and the proton core. The missing mass of the recoil nucleon is also calculated through the Monte Carlo program GENE2, according to the relation, $$W^2=M_p^2+2M_p(EE^{})4EE^{}\mathrm{sin}^2\frac{\theta }{2}$$ (29) The differential cross section $`d\sigma /d\mathrm{\Omega }`$ as a function of $`W^2`$ is compared in Fig.(8) with the experimental data of energies $`9.744,12.505`$ and $`15.730GeV`$. Fair agreement is obtained for the first reaction only. The deviation increases as the electron energy increases. This may be due to the fact that the model in hand has ignored the relative motion of the core inside the nucleon target. This relative motion let the recoil nucleon not has a unique value of missing mass, but instead, it posses a Gaussian -like distribution around $`W^2=M_p^2`$. The peak is wider at higher $`t`$. ## VI Conclusion 1- Data of elastic scattering may be reproduced in a wide range of the energy transferred using the (MPEM) with an expandable transition matrix representing ladder diagrams. 2- A transition matrix of order $`n`$, may be approximated as the power form of the one step ladder diagram at extremely high energy. 3- The power series of the transition matrix contains number of terms increasing with the energy transferred of the reaction. 4- Inelastic ep scattering are successfully described by the multi peripheral model. 5- The peripheral parameters are the dynamic parameters, which control the convergence of the particles in phase space. A critical value of peripheral parameter determines the critical surface at which phase transition may occur from the nuclear matter density to the quark gluon phase. 6- The rapidity distribution of the particles in the final state shows two peaks representing the forward and backward production. The peaks interfere together as the multiplicity increases in the final state.
warning/0001/hep-th0001186.html
ar5iv
text
# References Orbifold projection in supersymmetric QCD at $`N_fN_c`$ S. L. Dubovsky<sup>1</sup><sup>1</sup>1E-mail: sergd@ms2.inr.ac.ru— Institute for Nuclear Research of the Russian Academy of Sciences, 60th October Anniversary Prospect, 7a, 117312 Moscow, Russia ## Abstract Supersymmetric orbifold projection of $`𝒩=1`$ SQCD with relatively small number of flavors ($`N_fN_c`$) is considered. The purpose is to check whether orbifolding commutes with the infrared limit. On the one hand, one considers the orbifold projection of SQCD and obtains the low-energy description of the resulting theory. On the other hand, one starts with the low-energy effective theory of the original SQCD, and only then perfoms orbifolding. It is shown that at finite $`N_c`$ the two low-energy theories obtained in these ways are different. However, in the case of stabilized run-away vacuum these two theories are shown to coincide in the large $`N_c`$ limit. In the case of quantum modified moduli space, topological solitons carrying baryonic charges are present in the orbifolded low-energy theory. These solitons may restore the correspondence between the two theories provided that the soliton mass tends to zero in the large $`N_c`$ limit. 1. Recently, a new approach to the dynamics of strongly coupled gauge theories based on the correspondence between large $`N_c`$ theories and supergravity in higher dimensions has been suggested . One of impressive results obtained within this approach is a relation between Green’s functions of two different gauge theories in the large $`N_c`$ limit. The relation holds provided the gauge and matter contents are related by the so called “orbifold projection”. Then all Green’s functions of the projected (daughter) theory are equal to certain Green’s functions of the original (parent) theory. Suggested in the framework of string theory , this relation has been later supported by diagrammatic analysis at the level of field theory . By making use of supersymmetry breaking orbifold projection, several non-supersymmetric candidate dual pairs were suggested (see for string theoretical interpretation). In order to construct such a pair one starts with a supersymmetric model exhibiting the Seiberg duality . Then one makes orbifold projections of the electric and magnetic theories and arrives at two apparently different gauge models. The two models, however, are claimed to be equivalent in the large $`N_c`$ limit. A potential caveat of this construction is the necessity to interchange large $`N_c`$ and infrared limits. These limits may not commute if states with masses which scale as $`\mathrm{\Lambda }/N_c^\alpha `$, $`\alpha >0`$, are present in the spectrum ($`\mathrm{\Lambda }`$ is the infrared scale of the theory). Such states do not show up in the effective low-energy theory at finite $`N_c`$; however, they become massless in the large $`N_c`$ limit. This situation is inherent, e.g., in conventional QCD where $`\eta ^{}`$-meson becomes massless in the large $`N_c`$ limit. Another example is provided by $`𝒩=2`$ supersymmetric theories . Yet another potential problem is that the relation between parent and daughter theories proven at the level of planar diagrams may be spoiled by non-perturbative effects. Several arguments suggesting that these problems do not arise in $`𝒩=1`$ SQCD have been presented in Ref. . For instance, it was shown that large $`N_c`$ behavior of gluino condensate and of mesonic Green’s functions agrees with the exact results. Another possible check is to consider the situation where the low-energy descriptions of projected original and effective theories are known. Then one can explicitly check whether these two theories coincide in the large $`N_c`$ limit. One example of this type has been already presented in Ref. . Namely, supersymmetry preserving orbifold projection of SQCD in the region of Seiberg duality was considered there. This projection splits both the electric and magnetic theories into the sets of decoupled theories with smaller numbers of colors and flavors. These projected theories are again related to each other by the Seiberg duality, as expected. It is worth noting, that the equivalence between the projected theories holds even at finite $`N_c`$ in this case. The purpose of this letter is to study whether the equivalence of the above type holds in $`𝒩=1`$ SQCD with relatively small number of quark flavors, $`N_fN_c`$. In order to make use of the advantages of supersymmetry and keep the dynamics under control we restrict our consideration to orbifold projection that preserves supersymmetry. It is shown that at finite $`N_c`$ the equivalence between the projected theories does not hold. However, in the case of run-away vacuum stabilized by the quark mass term, the two theories are shown to coincide in the large $`N_c`$ limit (with $`N_f/N_c`$ kept constant). Consequently, this case serves as a non-trivial check of the commutativity of the large $`N_c`$ and infrared limits in $`𝒩=1`$ SQCD. In the case $`N_f=N_c`$, when quantum deformation of the classical flat directions occurs, the orbifolded effective theory does not reproduce the vacuum manifold of the orbifolded elementary theory at finite $`N_c`$. However, it is shown that topological solitons exist in this case. If the solitons become massless in the large $`N_c`$ limit, equivalence between two daughter theories may be restored. Finally, it is pointed out that the infrared limit and orbifold projection do not commute in the special case, when the parent theory belongs to the region of the Seiberg duality, while the daughter theory does not. This exception is due to the violation of the conditions of the theorem about orbifold projection in the parent magnetic theory. 2. Consider $`𝒩=1`$ SQCD with gauge group $`SU(\mathrm{\Gamma }N_c)`$ and $`\mathrm{\Gamma }N_f`$ flavors of quarks<sup>2</sup><sup>2</sup>2In what follows Latin and Greek letters stand for flavor and color indices, respectively. $`Q_\alpha ^a`$ and anti-quarks $`\stackrel{~}{Q}_b^\beta `$. $`\mathrm{\Gamma }`$ is some positive integer. By the large $`N_c`$ limit we mean the limit $`N_f,N_c\mathrm{}`$, with $`N_f/N_c`$ kept constant. A general description of the orbifold projection can be found in Refs. \[2–7\]. Let us describe its specific version adapted to the case of $`𝒩=1`$ SQCD which we consider throughout this paper. One makes use of the discrete group $`_\mathrm{\Gamma }`$. The action of the generator of this group on the quark superfields is defined as follows, $$Q_\alpha ^a\left(T_{N_f}\right)_b^a\left(T_{N_c}\right)_\alpha ^\beta Q_\beta ^b,$$ (1) where $`T_N`$ is a $`\mathrm{\Gamma }N\times \mathrm{\Gamma }N`$ diagonal matrix which consists of $`\mathrm{\Gamma }`$ blocks of the size $`N\times N`$, $$T_N=diag(1,e^{i/2\pi \mathrm{\Gamma }},\mathrm{},e^{(\mathrm{\Gamma }1)i/2\pi \mathrm{\Gamma }}).$$ The extension of this action to the anti-quark and gauge superfields is straightforward. The Lagrangian of the orbifolded theory is obtained from the Lagrangian of the parent theory by removing all fields and interactions which are not invariant under the action<sup>3</sup><sup>3</sup>3One can also consider the case when $`_\mathrm{\Gamma }`$ is non-trivially embedded into the non-anomalous $`R`$-symmetry group. However, supersymmetry is broken in the daughter theory in the latter case, so we will restrict our consideration to the case of trivial embedding. of $`_\mathrm{\Gamma }`$. The theorem proven in Refs. says that Green’s functions of the $`_\mathrm{\Gamma }`$–invariant fields calculated in the parent and daughter theories are the same at the level of planar diagrams modulo rescaling of coupling constants. Generally, in order to obtain the relation between parameters of the parent and daughter theories, one should rescale the fields in both theories in such a way that the corresponding Lagrangians take the form $$_p(g_{pi},fields)=\mathrm{\Gamma }N_cL_p(\overline{g}_{pi},fields)$$ (2) and $$_o(g_{oi},fields)=N_cL_o(\overline{g}_{oi},fields),$$ (3) where the subscripts $`p`$ and $`o`$ are assigned to parent and daughter (orbifolded) theories. Then the standard $`N_c`$-counting rules imply that couplings $`\overline{g}_{pi},\overline{g}_{oi}`$ are constant in the large $`N_c`$ limit. The equivalence between parent and daughter theories holds provided $`\overline{g}_{pi}`$ and $`\overline{g}_{oi}`$ are equal. In SQCD case under consideration this means that the canonical gauge coupling constants in the parent and orbifolded theories are related as follows, $$\mathrm{\Gamma }g_p^2=g_o^2.$$ (4) 3. Let us start from the case of run-away vacuum. The parent theory has $`\mathrm{\Gamma }N_c`$ colors and $`\mathrm{\Gamma }N_f`$ flavors with $`N_f/N_c<1`$. At low energies the dynamics can be described in terms of mesons $$M_b^a=Q_\alpha ^a\stackrel{~}{Q}_b^\alpha $$ (5) with the following effective superpotential $$W_{eff}=\mathrm{\Gamma }(N_cN_f)\left(\frac{\mathrm{\Lambda }_h^{\mathrm{\Gamma }(3N_cN_f)}}{detM}\right)^{\frac{1}{\mathrm{\Gamma }(N_cN_f)}},$$ (6) where $`\mathrm{\Lambda }_h`$ is the holomorphic infrared scale of the theory. The latter is related to the holomorphic coupling constant $`g_h`$ in the following way, $$\mathrm{\Lambda }_h^{3N_cN_f}=\mu e^{8\pi ^2/g_h^2(\mu )},$$ where $`\mu `$ is the normalization scale. The value of the holomorphic coupling constant $`g_h`$ is determined by the Shifman-Vainstein relation between the holomorphic and canonical coupling constants, $$Re\left(\frac{8\pi ^2}{g_h^2}\right)=\frac{8\pi ^2}{g^2}+N_c\mathrm{ln}g^2.$$ (7) Besides mesons, the low-energy effective theory contains pure Yang-Mills sector corresponding to the unbroken $`SU(\mathrm{\Gamma }(N_cN_f))`$ gauge group. In order to protect mesons from acquiring infinite vacuum expectation values, we consider a deformation of the theory by the mass term $`mQ^a\overline{Q}_a=m\text{Tr}M`$ added to the superpotential. We take this mass term in the flavor-symmetric form to preserve the vectorial $`SU(\mathrm{\Gamma }N_f)`$ symmetry relevant to $`1/N_c`$ expansion. Now we apply the orbifold projection to the high-energy theory and study whether the obtained daughter theory corresponds at low-energies to the orbifolded effective theory. Upon orbifolding the high-energy theory one obtains $`\mathrm{\Gamma }`$ decoupled theories. Each of them has the gauge group $`SU(N_c)`$ and $`N_f`$ quark flavors. The rescaling rule (4) implies that holomorphic infrared scale of the orbifolded theory is equal to $$\mathrm{\Lambda }_o=\mathrm{\Gamma }^{\frac{N_c}{3N_cN_f}}\mathrm{\Lambda }_h.$$ (8) The quark mass is left unchanged under the orbifold projection. Consequently, at low energies the daughter theory is described by the following effective superpotential, $$W_1=\underset{i=1}{\overset{\mathrm{\Gamma }}{}}(N_cN_f)\left(\frac{\mathrm{\Lambda }_h^{3N_cN_f}}{\mathrm{\Gamma }^{N_c}detM^{(i)}}\right)^{\frac{1}{N_cN_f}}+\underset{i=1}{\overset{\mathrm{\Gamma }}{}}mTrM^{(i)},$$ (9) where mesons referring to different gauge sectors are distinguished by the superscript $`(i)`$. In addition there is a pure gluonic sector described by $`\left[SU(N_cN_f)\right]^\mathrm{\Gamma }`$ gauge group. Let us now consider the orbifold projection of the effective theory, the latter consisting of the mesonic sector described by the superpotential (6) and of the gluonic sector described by the gauge group $`SU(\mathrm{\Gamma }(N_cN_f))`$. The action of the discrete group $`_\mathrm{\Gamma }`$ on the meson superfields is determined by Eqs. (1) and (5). In the effective meson theory, $`\mathrm{\Lambda }_h`$ serves as the coupling constant. Consequently, one rescales it according to the general rule described above. The weak coupling form of the mesonic Kahler potential, $`K(M^{},M)=2Tr(M^{}M)^{1/2}`$ (or, equivalently, the explicit relation between the meson and quark superfields (5)), implies that rescaling $`M\mathrm{\Gamma }N_cM`$ is needed to cast the Lagrangian of the effective theory in the form (2). Then, the general rescaling rule described above implies that $`\mathrm{\Lambda }_h`$ again rescales according to Eq. (8). After the orbifold projection of the low-energy theory one obtains the gluonic sector described by the same $`\left[SU(N_cN_f)\right]^\mathrm{\Gamma }`$ gauge group as above and the following meson superpotential, $$W_2=\mathrm{\Gamma }(N_cN_f)\underset{i=1}{\overset{\mathrm{\Gamma }}{}}\left(\frac{\mathrm{\Lambda }_h^{3N_cN_f}}{\mathrm{\Gamma }^{N_c}detM^{(i)}}\right)^{\frac{1}{\mathrm{\Gamma }(N_cN_f)}}+\underset{i=1}{\overset{\mathrm{\Gamma }}{}}mTrM^{(i)}.$$ (10) Obviously, the two superpotentials (9) and (10) are different. However, the vacuum expectation values of mesons are the same in both cases, $$M^{(i)}=\mathrm{\Gamma }^1\mathrm{\Lambda }_h^{\frac{3N_cN_f}{N_c}}m^{\frac{N_fN_c}{N_c}}M.$$ (11) Moreover, one can straightforwardly check that the Lagrangians describing dynamics in these vacua are the same for both superpotentials in the large $`N_c`$ limit. To see this, let us compare $`F`$–terms originating from Eqs. (9) and (10), which determine the scalar potentials in the two cases. Making use of the first superpotential (9) we obtain $$F_1^{(i)}\frac{W_1}{M^{(i)}}=\left(\frac{\mathrm{\Lambda }_h^{3N_cN_f}}{\mathrm{\Gamma }^{N_c}detM^{(i)}}\right)^{\frac{1}{N_cN_f}}\left(M^{(i)T}\right)^1+m.$$ In the case of the second superpotential (10) we have $$F_2^{(i)}\frac{W_2}{M^{(i)}}=\left(\underset{j=1}{\overset{\mathrm{\Gamma }}{}}\frac{\mathrm{\Lambda }_h^{3N_cN_f}}{\mathrm{\Gamma }^{N_c}detM^{(j)}}\right)^{\frac{1}{\mathrm{\Gamma }(N_cN_f)}}\left(M^{(i)T}\right)^1+m.$$ We write the meson fields in the form $`M^{(i)}=M+\delta M^{(i)}`$, then $`F_1^{(i)}`$ and $`F_2^{(i)}`$ differ by terms coming from the expansion of determinants in the denominators. However, these terms are suppressed by extra powers of $`N_cN_f`$. Hence, both $`F`$–terms in the large $`N_c`$ limit are reduced to $$F^{(i)}=\mathrm{\Gamma }^1m^{\frac{N_f}{N_c}}\mathrm{\Lambda }^{\frac{3N_cN_f}{N_c}}\left(M+\delta M^{(i)T}\right)^1+m.$$ We conclude that $`𝒩=1`$ SQCD with $`N_f<N_c`$ provides a non-trivial check of the technique suggested in Ref. . 4. Let us now turn to the case of SQCD with the gauge group $`SU(\mathrm{\Gamma }N_c)`$ and $`\mathrm{\Gamma }N_c`$ quark flavors. This theory exhibits quantum deformation of the moduli space . Namely, the space of vacua of the microscopic theory is described by the set of holomorphic gauge invariants constructed of quark and anti-quark fields. These invariants are mesons (5), and (anti)-baryons $$B=ϵ^{\alpha _1..\alpha _{\mathrm{\Gamma }N_c}}Q_{\alpha _1}^1..Q_{\alpha _{\mathrm{\Gamma }N_c}}^{\mathrm{\Gamma }N_c},$$ $$\stackrel{~}{B}=ϵ_{\beta _1..\beta _{\mathrm{\Gamma }N_c}}\stackrel{~}{Q}_1^{\beta _1}..\stackrel{~}{Q}_{\mathrm{\Gamma }N_c}^{\beta _{\mathrm{\Gamma }N_c}},$$ subject to the constraint $$detMB\stackrel{~}{B}=\mathrm{\Lambda }_h^{2\mathrm{\Gamma }N_c}.$$ (12) The r.h.s of Eq. (12) is of purely quantum origin and indicates the difference between the topologies of the quantum and classical spaces of vacua. At low energies this theory is described by the non-linear sigma model with the field space parameterized by mesonic $`M_{\stackrel{~}{b}}^a`$ and (anti-)baryonic $`(\stackrel{~}{B})B`$ coordinates that satisfy the constraint (12). Let us take the orbifold projection of this theory with respect to the discrete group $`_\mathrm{\Gamma }`$ as described above. Then, in analogy to the case of smaller number of flavors, the orbifolded theory splits into $`\mathrm{\Gamma }`$ copies of SQCD with gauge groups $`SU(N_c)`$ and $`N_c`$ quark flavors. At low energy the latter theory is described in terms of sigma-model fields $`M_b^{a(i)}`$, $`B^{(i)},\stackrel{~}{B}^{(i)}`$, where $`a,b=1,\mathrm{},N_c`$ and $`i=1,\mathrm{},\mathrm{\Gamma }`$. These fields satisfy the constraints $$detM^{(i)}B^{(i)}\stackrel{~}{B}^{(i)}=\mathrm{\Lambda }_o^{2N_c}$$ (13) for every $`i`$. The orbifold projection of the low-energy theory described by Eq. (12) is the non-linear sigma-model with the fields $`M_b^{a(i)}`$, $`B,\stackrel{~}{B}`$ subject to the constraint $$\underset{i=1}{\overset{\mathrm{\Gamma }}{}}detM^{(i)}B\stackrel{~}{B}=\mathrm{\Lambda }_o^{2\mathrm{\Gamma }N_c}.$$ (14) Manifolds $`𝒬_1`$ and $`𝒬_2`$ determined by Eqs. (13) and (14), respectively, are different. For instance, the former has complex dimension $`\mathrm{\Gamma }(N_c^2+1)`$ whereas the complex dimension of $`𝒬_2`$ is $`(\mathrm{\Gamma }N_c^2+1)`$. Consequently, the orbifolded original theory, described at low energies by Eq. (13), has extra $`(\mathrm{\Gamma }1)`$ massless superfields as compared to the orbifolded sigma model. This is due to the fact that baryons of the parent sigma model do not carry flavor or color indices and, as a result, they are not affected by the orbifold projection. Hence, the orbifolded effective theory has one conserved baryonic charge and one pair of massless baryons instead of $`\mathrm{\Gamma }`$. Furthermore, there is a profound difference between $`𝒬_1`$ and $`𝒬_2`$. To see this difference, let us study their topology in more detail. The manifold $`𝒬_1`$ is merely a direct product of $`\mathrm{\Gamma }`$ copies of the manifold $`𝒬`$ defined by one of the equations listed in Eqs. (13). The topology of $`𝒬`$ has been studied in Ref. . It was found there that this space is homotopically equivalent to the double suspension of $`SL(N_c,)`$ group, $`\mathrm{\Sigma }(\mathrm{\Sigma }(SL(N_c,)))`$. The suspension of the manifold $`𝒳`$ is the cylinder $`𝒳\times [0,1]`$ with all points on the lower base $`𝒳\times 0`$ identified and all points on the upper base $`𝒳\times 1`$ identified as well. For example, the suspension of $`d`$-dimensional sphere $`S^d`$ is $`(d+1)`$-dimensional sphere $`S^{(d+1)}`$. The latter observation is the basis of the theorem about suspension (see., e.g., Ref. , p.79), $`\pi _{q+1}\left(\mathrm{\Sigma }\left(𝒳\right)\right)=\pi _q\left(𝒳\right)`$ for all $`q2n2`$, provided that $`\pi _i(𝒳)=0`$ for $`i<n`$. In particular, this theorem implies that the lowest non-trivial homotopy group of $`𝒬`$ is $`\pi _5(𝒬)=`$, so that $`\pi _5(𝒬_1)=^\mathrm{\Gamma }`$ . Let us now consider the manifold $`𝒬_2`$. A straightforward generalization of the arguments of Ref. shows that this space is homotopically equivalent to the double suspension of the manifold $`𝒴`$ determined by the following equation, $$\underset{i=1}{\overset{\mathrm{\Gamma }}{}}detM^{(i)}=\mathrm{\Lambda }_o^{2\mathrm{\Gamma }N_c}.$$ (15) At $`\mathrm{\Gamma }=1`$, the manifold $`𝒴`$ is $`SL(N_c,)`$ in accordance with the above discussion. Let us calculate the lowest homotopy group of $`𝒬_2`$. Every non-degenerate matrix $`M^{(i)}`$ can be decomposed into a product of a matrix with unit determinant and of the diagonal matrix of the form $`diag(t_i,1,\mathrm{},1)`$. Then Eq. (15) implies that the manifold $`𝒴`$ is a direct product $`SL(N_c,)^\mathrm{\Gamma }\times T_{\mathrm{\Gamma }1}`$. Here $`T_{\mathrm{\Gamma }1}`$ is $`(\mathrm{\Gamma }1)`$-dimensional complex torus $`_{}^{\mathrm{\Gamma }1}`$ determined by the equation $$\underset{i=1}{\overset{\mathrm{\Gamma }}{}}t_i=\mathrm{\Lambda }_o^{2\mathrm{\Gamma }N_c}.$$ Therefore, the manifold $`𝒬_2`$ is homotopically equivalent to $$𝒬_2\mathrm{\Sigma }(\mathrm{\Sigma }(𝒴))𝒬_1\times (S^3)^{\mathrm{\Gamma }1}.$$ The lowest non-trivial homotopy group of $`𝒬_2`$ is $$\pi _3(𝒬_2)=^{\mathrm{\Gamma }1}.$$ (16) As a result, contrary to the case of $`𝒬_1`$ sigma-model, there may exist topological solitons in the $`𝒬_2`$ sigma-model, provided that the stabilizing higher-derivative terms are present in the Kahler potential of the effective theory. These solitons are distinguished by $`(\mathrm{\Gamma }1)`$ conserved topological charges. It is tempting to identify these charges with the missing $`(\mathrm{\Gamma }1)`$ baryonic charges. An argument in favor of this identification is the possibility to construct corresponding topological currents by formal extension of the algebra of the Noether currents; in analogy to the case of the conventional QCD . This is possible due to the existence of the Wess–Zumino term in SQCD with $`N_f=N_c`$. In the parent theory this term reads as follows , $$\mathrm{\Gamma }_p=\frac{1}{12\pi ^2\mathrm{\Lambda }^{4\mathrm{\Gamma }N_c}}Im𝑑\mathrm{\Omega }detMϵ^{\mu \nu \lambda \rho \sigma }_\mu B_\nu \stackrel{~}{B}\times Tr\left(M^1_\lambda MM^1_\rho MM^1_\sigma M\right).$$ After the orbifold projection it takes the following form, $`\mathrm{\Gamma }_o={\displaystyle \frac{1}{12\pi ^2\mathrm{\Lambda }_o^{4\mathrm{\Gamma }N_c}}}Im{\displaystyle 𝑑\mathrm{\Omega }\left(\underset{i=1}{\overset{\mathrm{\Gamma }}{}}detM^{(i)}\right)ϵ^{\mu \nu \lambda \rho \sigma }_\mu B_\nu \stackrel{~}{B}}`$ $`\times {\displaystyle \underset{i=1}{\overset{\mathrm{\Gamma }}{}}}Tr\left(M^{(i)1}_\lambda M^{(i)}M^{(i)1}_\rho M^{(i)}M^{(i)1}_\sigma M^{(i)}\right).`$ The contribution of this term to the flavor current in the daughter theory reads as follows, $$j^\mu =\frac{1}{4\pi ^2\mathrm{\Lambda }_o^{4N_c}}\underset{i=1}{\overset{\mathrm{\Gamma }}{}}detM^{(i)}ϵ^{\mu \nu \lambda \rho }_\nu B_\lambda \stackrel{~}{B}Tr\left(T^f\underset{i=1}{\overset{\mathrm{\Gamma }}{}}_\rho M^{(i)}M^{(i)1}\right)+h.c.,$$ where $`T^f`$ is a generator of the flavor group. In complete analogy to the case of QCD, one formally substitutes here $`T^f`$ in the form $`diag(0,\mathrm{},i,\mathrm{},0)`$ where $`0,i`$ stand for the blocks of length $`\mathrm{\Gamma }`$. In this way one obtains $`\mathrm{\Gamma }`$ conserved topological currents, $$j^{(i)\mu }=\frac{1}{2\pi ^2\mathrm{\Lambda }_o^{4N_c}}Im\underset{i=1}{\overset{\mathrm{\Gamma }}{}}detM^{(i)}ϵ^{\mu \nu \lambda \rho }_\nu B_\lambda \stackrel{~}{B}Tr\left(_\rho M^{(i)}M^{(i)1}\right),$$ which are subject to the constraint $`j^{(i)\mu }=0`$. These currents correspond precisely to the topological property (16), and the above argument indeed shows that they are naturally identified with the missing baryonic currents. To prove the equivalence of the $`𝒬_1`$ and $`𝒬_2`$ sigma-models in the large $`N_c`$ limit one would have to study the behavior of the soliton mass in this limit. The above consideration indicates that these two models may be equivalent at large $`N_c`$ provided that the soliton mass tends to zero as $`N_c\mathrm{}`$. We leave the analysis of this point for future. 5. Finally, let us present an example in which the orbifold projection does not commute with the infrared limit. Let us consider $`𝒩=1`$ SQCD with $`\mathrm{\Gamma }N_c`$ colors and $`\mathrm{\Gamma }(N_c+1)`$ flavors. This theory belongs to the region of the Seiberg duality and its low-energy behavior can be described in terms of a magnetic theory with $`\mathrm{\Gamma }`$ colors and $`\mathrm{\Gamma }(N_c+1)`$ flavors. In addition, the magnetic theory contains gauge-singlet meson fields $`M_b^a`$ with the following superpotential $$W_m=qM\stackrel{~}{q},$$ where $`q`$ and $`\stackrel{~}{q}`$ are dual quark and anti-quark superfields; a dimensionful coefficient is omitted. Upon orbifolding the electric theory one obtains $`\mathrm{\Gamma }`$ decoupled theories with the gauge group $`SU(N_c)`$ and $`(N_c+1)`$ quark flavors. At low energies this theory confines and describes mesons and baryons interacting through the following superpotential $$W_1=\underset{i=1}{\overset{\mathrm{\Gamma }}{}}\left(B^{(i)}M^{(i)}\stackrel{~}{B}^{(i)}detM^{(i)}\right).$$ (17) The orbifold projection of the magnetic theory, on the other hand, leaves $`\mathrm{\Gamma }`$ pairs of the gauge-singlet superfields $`q^{(i)}`$ and $`\stackrel{~}{q}^{(i)}`$ with the same quantum numbers as $`B^{(i)}`$ and $`\stackrel{~}{B}^{(i)}`$ and splits the meson multiplet in the same manner as above. The projected superpotential reads as follows, $$W_2=\underset{i=1}{\overset{\mathrm{\Gamma }}{}}q^{(i)}M^{(i)}\stackrel{~}{q}^{(i)},$$ (18) which is different from Eq. (17). The reason of this discrepancy is that $`1/N_c`$ expansion does not work in the magnetic theory, as the number of magnetic colors does not depend on $`N_c`$. Consequently, the relation between the planar diagrams in the parent and daughter theories does not lead to the relation between corresponding Green’s functions in this case. 6. To conclude, we have considered orbifold projection of SQCD with relatively small number of quark flavors. It was found that in the case of stabilized run-away vacuum, the orbifold projection serves as a non-trivial check of the commutativity of the large $`N_c`$ and infrared limits. We discussed in sect. 5 also a specific way of taking $`N_c\mathrm{}`$ such that two daughter theories obtained in different limits are not equivalent. The reason is that the conditions of the theorem about the orbifold projection are violated in the effective theory. The most intriguing situation occurs in SQCD with quantum modified moduli space, $`N_f=N_c`$. In this case it was found that upon orbifolding the high-energy and low-energy theories, one obtains a pair of sigma models which are not equivalent to each other at finite $`N_c`$. Namely, the vacuum manifold of the orbifolded elementary theory has larger dimension than the space of vacua of the orbifolded effective theory. As a result, the latter has smaller number of massless fields in every particular vacuum. Moreover, the number of the Noether currents in the orbifolded effective theory is too small to reproduce the correct current algebra of the sigma model corresponding to orbifolded SQCD. However, the two sigma models may become equivalent at large $`N_c`$ due to the presence of topological solitons in the orbifolded effective theory. Namely, the topological currents restore the structure of the current algebra, and the solitons may provide the correspondence between the spectra of light fields in the two theories provided that the soliton mass tends to zero at large $`N_c`$. The correct structure of the space of vacua may be restored due to the presence of topologically non-trivial field configurations corresponding to non-zero “soliton condensate”. Further analysis of this model from both field theoretical and brane points of view may provide new insights into orbifold projection. The author acknowledges numerous helpful discussions with F.L. Bezrukov, D.S. Gorbunov, A.A. Penin and V.A. Rubakov. This work is supported in part by Russian Foundation for Basic Research grant 99-02-18410, by the Russian Academy of Sciences, JRP grant 37 and by ISSEP fellowship.
warning/0001/cond-mat0001244.html
ar5iv
text
# Far-infrared reflectivity of Bi2Sr2CuO6: the “anomalous Drude” model and the optical pseudogap revisited ## Abstract An optical “pseudogap” is assumed to open at low $`T`$ in the “anomalous Drude” absorption, which models the optical conductivity $`\sigma (\omega )\omega ^1`$ of high-$`T_c`$ superconductors by a linewidth $`\mathrm{\Gamma }10^3`$ cm<sup>-1</sup> varying with $`\omega `$. In the $`\sigma (\omega )`$ of Bi<sub>2</sub>Sr<sub>2</sub>CuO<sub>6</sub> measured down to 10 cm<sup>-1</sup>, we have resolved, instead, two components separated by a deep minimum: i) a normal Drude term with $`\mathrm{\Gamma }`$=35 cm<sup>-1</sup> at 30 K, in very good agreement with transport data; ii) a strong band peaked in the far infrared (FIR), likely due to bound charges, whose tail exhibits the $`\omega ^1`$ dependence. As the FIR peak softens for $`T0`$, it opens a pseudogap-like depression in $`\sigma (\omega )`$ accordingly to ordinary sum rules. Several studies have been published during recent years on the possible implications for high-$`T_c`$ superconductivity of the “pseudogap”. This term indicates a depression in the low-energy continuum of states of many high-$`T_c`$ superconductors (HCTS), that has been observed by different techniques below a characteristic temperature $`T^{}>>T_c`$. The pseudogap has been observed also in the optical spectra of a number of metallic cuprates. These include underdoped systems like Bi<sub>2</sub>Sr<sub>2</sub>CuO<sub>6</sub> with $`T_c`$ = 8 K and HgBa<sub>2</sub>Ca<sub>2</sub>Cu<sub>3</sub>O<sub>8+y</sub> with $`T_c`$ = 121 K, overdoped cuprates like La<sub>2-x</sub>Sr<sub>x</sub>CuO<sub>4</sub> with $`x>0.18`$ and Tl<sub>2</sub>Ba<sub>2</sub>CuO<sub>6+y</sub>, and even metals close to optimum doping like HgBa<sub>2</sub>Ca<sub>2</sub>Cu<sub>3</sub>O<sub>8+y</sub> with $`T_c`$ = 130 K and Bi<sub>2</sub>Sr<sub>2</sub>CaCu<sub>2</sub>O<sub>8+y</sub> with $`T_c`$ = 90 K. In these experiments, the free-carrier contribution to the real part $`\sigma (\omega )`$ of the optical conductivity is described by an “anomalous Drude” term $$\sigma (\omega )=\frac{\omega _D^2/4\pi }{\mathrm{\Gamma }^{}(\omega )+[m^{}(\omega )/m]^2\omega ^2/\mathrm{\Gamma }^{}(\omega )},$$ (1) where $`\omega _D`$ is a constant plasma frequency, while both the scattering rate $`\mathrm{\Gamma }^{}`$ and the effective mass $`m^{}`$ depend on the photon energy $`\omega `$. This model has been originally introduced to fit the $`\sigma (\omega )\omega ^1`$ law which replaces the $`\omega ^2`$ behavior of conventional metals in the mid-infrared spectra of HCTS. The use of Eq. (1) leads to large scattering rates for the carriers in the cuprates. At $`TT^{}`$, $`\mathrm{\Gamma }^{}1500`$ cm<sup>-1</sup> in the far infrared, for both La<sub>2-x</sub>Sr<sub>x</sub>CuO<sub>4</sub> and HgBa<sub>2</sub>Ca<sub>2</sub>Cu<sub>3</sub>O<sub>8+y</sub>. Below $`T^{}`$, $`\mathrm{\Gamma }^{}`$ decreases to $``$ 500 cm<sup>-1</sup> in the former compound, to $``$ 1000 cm<sup>-1</sup> in the latter. Thus, even if the opening of an optical pseudogap below $`T^{}`$ is displayed by $`\sigma (\omega )`$, its amplitude is generally extracted from the above drop in the far-infrared (FIR) part of $`\mathrm{\Gamma }^{}(\omega )`$. The idea of an optical pseudogap is therefore intrinsically related to the anomalous Drude approach of Eq. (1) and to the assumption that the mid-infrared absorption of cuprates is dominated by some anomalous Fermi liquid. However, the above picture leads to some inconsistencies. For instance, in Bi<sub>2</sub>Sr<sub>2</sub>CaCu<sub>2</sub>O<sub>8+y</sub> samples with similar $`T_c`$’s, the pseudogap determined optically ($`700`$ cm<sup>-1</sup>) is much larger than that found in angle-resolved photoemission (0-260 cm<sup>-1</sup>, depending on the direction in the $`k`$\- space). This discrepancy can hardly be explained, if one also considers that the optical absorption results from an average on the Fermi surface. Moreover, an optical pseudogap is observed in La<sub>2-x</sub>Sr<sub>x</sub>CuO<sub>4</sub>, although no such effect appears in the spin-lattice relaxation rate of this cuprate. Finally, in the metallic spectra of Refs. $`\sigma (\omega )`$ surprisingly decreases for $`\omega `$ 0, for any $`T`$ of the normal phase. This decrease is observed on the low-frequency side of a strong FIR peak which is systematically associated with the observation of the pseudogap. However, the FIR peak is attributed to scattering of the free carriers by oxygen atoms randomly distributed in the Cu-O planes. As this explanation does not seem to be related to the existing theoretical models of the pseudogap, the link between this latter and the FIR peak remains obscure. A first hint comes from the optical conductivity of Nd<sub>1.88</sub>Ce<sub>0.12</sub>CuO<sub>4</sub> (NCCO), partially reported in Ref. . A peak develops in the infrared, which softens considerably as $`T`$ is lowered. This shift causes at higher energies a depression in $`\sigma (\omega )`$ around 1000 cm<sup>-1</sup>, very similar to the pseudogap reported by the authors cited above. This suggests that, in NCCO, the optical pseudogap opens by a transfer of spectral weight to the FIR peak as this softens and narrows, as requested by ordinary sum rules. One should then focus on this peak at finite frequency, attributed in NCCO to charges bound to the lattice via a polaronic coupling. However, Nd<sub>1.88</sub>Ce<sub>0.12</sub>CuO<sub>4</sub> is electron doped and semiconducting, even if very close to the insulator-to-metal transition. In order to study in greater detail a system where the “anomalous-Drude + pseudogap” model has already been applied and with the lowest carrier density, we have measured the reflectivity of Bi<sub>2</sub>Sr<sub>2</sub>CuO<sub>6</sub> (BSCO), from 400 to 8 K and from 15,000 to 10 cm<sup>-1</sup>. By extending the measurements in BSCO to the very-far-infrared region of frequencies, we have first resolved from the FIR peak a normal Drude contribution which is in agreement with the dc conductivity, $`\sigma _{dc}`$, and the London penetration depth of the material. Basing on these results we show that, at least in BSCO: i) $`\sigma (\omega ,T)`$ cannot be explained in terms of a one-component model like the anomalous Drude of Eq. (1); ii) the pseudogap does not open in the continuum of states of some kind of (anomalous) free carrier. On the contrary, we show that: i) only the coexistence of two type of carriers may account for the optical absorption; ii) the optical pseudogap is an effect created by the temperature-dependent absorption of some weakly bound charges. In such context, the phase-separation model proposed by Emery and Kivelson is found to be in good agreement with the present observations. The sample investigated here is a thick BSCO film, highly oriented with the $`c`$-axis orthogonal to the surface. It has been grown by liquid phase epitaxy on a LaGaO<sub>3</sub> substrate. Its resistivity $`\rho (T)`$, obtained from standard 4-points measurements, is reported by a full line in the inset of Fig. 1. It shows a linear dependence on $`T`$ from 300 to about 65 K, where a slight change of slope is observed. The superconducting transition has its onset at $`T_c`$ = 20 K, with a width of 5 K. The $`\rho `$ value at 300 K (1.4 $`\times 10^3\mathrm{\Omega }`$cm) is intermediate between that of a good BSCO single crystal (0.3 $`\times 10^3\mathrm{\Omega }`$cm) and that of a polycrystalline pellet of the same material (2.7 $`\times 10^3\mathrm{\Omega }`$cm). This may be attributed to the presence of grain boundaries in the well oriented $`ab`$ plane of the present film. However, eventual grain boundaries are not expected to affect the infrared measurements. Indeed, the reflectivity $`R(\omega )`$ of the BSCO film, relative to a gold-plated reference and reported in Fig. 1, is typical of the best single crystals. The $`R(\omega )`$ spectra, with the electric field polarized in the $`ab`$ plane, were collected using a Bomem DA8 interferometer coupled to the infrared synchrotron radiation beamline SIRLOIN of the LURE laboratory at Orsay. Helium-cooled bolometers, mercury-cadmium-tellurium or silicon detectors were used, depending on the frequency range under investigation. The large size of the film (0.5x0.5 cm<sup>2</sup>) and the good performances of the apparatus have allowed us to measure $`R(\omega )`$ down to unusually low values of $`\omega `$ ($``$ 10 cm<sup>-1</sup>). The film thickness (1.8 $`\mu `$m) was such that no correction for the substrate contribution to $`R(\omega )`$ was needed; see Fig. 1. Therefore, the optical conductivity $`\sigma (\omega )`$ was obtained by simple, canonical Kramers-Kronig transformations of $`R(\omega )`$. In the normal metallic phase ($`T>T_c`$), $`R(\omega )`$ has been extrapolated from $`\omega `$ = 10 cm<sup>-1</sup> to $`\omega `$ = 0 by a Drude-Lorentz fit, in the superconducting phase by a London conductivity, as usually done; see Ref. . Moreover, $`R(\omega )`$ has been extrapolated from 15,000 cm<sup>-1</sup> up to 320,000 cm<sup>-1</sup> by using the data of Ref., that have been extended to higher frequencies by a $`\omega ^4`$ law. The conductivities $`\sigma (\omega ,T)`$ corresponding to the raw reflectivities of Fig. 1 are shown in Fig. 2. At any $`T`$, $`\sigma (\omega ,T)`$ exhibits a broad band peaked in the far infrared (FIR peak). Its high-frequency side behaves as $`\omega ^1`$ up to 6000 cm<sup>-1</sup>, as shown in the inset of Fig. 2 for $`T`$ = 200 and 30 K. The tail of the FIR peak accounts, therefore, for the frequency dependence of the so-called anomalous-free-carrier absorption usually observed in metallic cuprates. The peak frequency ($``$ 800 cm<sup>-1</sup> at 400 K) increasingly softens and narrows as $`T`$ is lowered, until it reaches 110 cm<sup>-1</sup> at 30 K. In the mid infrared, a broad pseudogap opens for $`T0`$. This feature is less deep than that observed in other cuprates (see Refs. ), most likely because the doping of the present BSCO sample is nearly optimal. This pseudogap is just due to the red shift and the narrowing of the FIR peak for decreasing $`T`$, which produce a transfer of spectral weight towards low frequencies. Indeed, the effective number of carriers per unit cell $`n_{eff}`$ is constant with $`T`$ within a few percent, for all temperatures of the normal phase, see the inset in Fig. 3. $`n_{eff}`$ has been evaluated from the relation $$n_{eff}=\frac{2m^{}V}{\pi e^2}_{\omega _{min}}^{\omega _{max}}\sigma (\omega )𝑑\omega ,$$ (2) where $`V`$ is the cell volume, $`\omega _{min}`$ = 10 cm<sup>-1</sup>, $`\omega _{max}`$ = 7500 cm<sup>-1</sup>, and $`m^{}`$ has been assumed equal to the free electron mass. The results of Fig. 2 are quite similar to those reported in Ref. for a single crystal of BSCO with $`T_c`$ = 8 K, whose $`R(\omega )`$ was measured from 12,000 to 50 cm<sup>-1</sup>. However, as evident from Fig. 3 where the FIR $`\sigma (\omega ,T)`$ has been reported on an expanded scale, the extension of the spectra down to 10 cm<sup>-1</sup> has allowed us to resolve a narrow, low-frequency absorption from the FIR peak. The dip separating these two components of $`\sigma (\omega )`$ is observed at all temperatures. It corresponds to a change of slope at $``$ 50 cm<sup>-1</sup> in $`R(\omega )`$, see Fig. 1, where no instrumental effects are present, e. g., beamsplitter transmittance minima or data file overlaps. The low-frequency component of the absorption increases for $`\omega 0`$, extrapolates to the $`\sigma _{dc}(T)`$ values measured in the same sample (full triangles), and disappears below $`T_c`$. The open circles in Fig. 3 show that this feature, presumably peaked at $`\omega =0`$, is very well fitted at 30 K by a normal Drude term $$\sigma (\omega )=\frac{\omega _D^2/4\pi }{\mathrm{\Gamma }_D+\omega ^2/\mathrm{\Gamma }_D},$$ (3) with plasma frequency $`\omega _D`$ = 2100 cm<sup>-1</sup> and scattering rate $`\mathrm{\Gamma }_D`$ = 35 cm<sup>-1</sup>. At 100 (200) K one obtains, instead, $`\omega _D`$ = 1800 (1600) cm<sup>-1</sup> and $`\mathrm{\Gamma }_D`$ = 35 (40) cm<sup>-1</sup>. If one extrapolates the Drude fits to obtain $`\sigma (0,T)`$ at all temperatures, one obtains the values reported by open circles in the inset of Fig. 1 together with the dc resistivity (full line) measured in the same sample. The agreement is within a few percent. The results reported in Figs. 2 and 3 point towards a multi-component model for $`\sigma (\omega )`$ and indicate a coexistence of free and bound charges. The latters may be polaronic in nature and may aggregate at low temperatures, as proposed for the NCCO system in Ref. . Emery and Kivelson have presented a picture where free carriers are scattered by arrays of (dynamical) bound charges which carry local dipoles. Following these authors, the optical conductivity can be written as $$\sigma (\omega ,T)=\sigma _a+\sigma _b=\frac{e^2A}{\omega }\chi _2(\omega ,T)+(e^{})^2\omega \chi _2(\omega ,T)$$ (4) with $$\chi _2(\omega ,T)=c\mathrm{tanh}(\mathrm{}\omega /2k_BT)\frac{\mathrm{\Gamma }_p}{\mathrm{\Gamma }_p^2+\omega ^2}.$$ (5) In Eq. (4), $`\sigma _b`$ is the contribution of the bound charges, modeled as $`c`$ dipoles of charge $`e^{}`$. $`\sigma _a`$ takes into account the scattering of the free carriers of charge $`e`$ by these dipoles. $`\mathrm{\Gamma }_p`$ determines both the FIR peak frequency and width, while $`A`$ is a constant whose expression is reported in Ref. . Equation (4) was compared unsuccessfully with the optical conductivity measured in La<sub>2</sub>CuO<sub>4+y</sub>. On the contrary, good fits of Eq. (4) to the present BSCO data are obtained at all temperatures, as reported in Fig. 3 for $`T>T_c`$ (200 K) and $`T<T_c`$ (8 K). Remarkably, by using the two parameters $`c`$ and $`\mathrm{\Gamma }_p`$ (whose values are 230 cm<sup>-1</sup> at 200 K, 135 cm<sup>-1</sup> at 8 K), one reproduces the Drude term, the asymmetric FIR peak, and the softening and narrowing of this latter for decreasing $`T`$. Below $`T_c`$ the Drude peak disappears from the measuring range and the FIR peak looses the intensity corresponding to the underlying Drude tail. This effect, which has been reproduced in the fit by fixing $`\sigma _a(\omega ,T=8K)`$ = 0, allows one to perform a further check of the present analysis. In fact, the loss of spectral weight observed below $`T_c`$ in Fig. 3 provides an estimate of the London penetration depth in the film, as given by the Ferrell-Glover sum rule $$\lambda _L^2=\frac{c^2}{8_{\omega _{min}}^{\omega _{max}}[\sigma _n(\omega ,30K)\sigma _s(\omega ,8K)]𝑑\omega },$$ (6) where here $`c`$ is the speed of light and the other symbols have a obvious meaning. For $`\omega _{min}`$ = 10 cm<sup>-1</sup> and $`\omega _{max}`$ = 15,000 cm<sup>-1</sup> one obtains $`\lambda _L=300\pm 10`$ nm, a value in excellent agreement with that obtained for Bi<sub>2</sub>Sr<sub>2</sub>CuO<sub>6</sub> from transport measurements (310 nm). On the other hand, no estimate of the superconducting gap can be made, due to the strong FIR peak which overshadows the collapse of the Drude component below $`T_c`$. In conclusion, the present far-infrared study of BSCO questions the widely accepted “anomalous Drude + pseudogap” model, through the identification in $`\sigma (\omega ,T)`$ of a narrow, normal Drude absorption well resolved from a broad FIR peak. The small value of the free-carrier scattering rate ($`\mathrm{\Gamma }_D<`$ 50 cm<sup>-1</sup>) is consistent with a metal in an extremely clean limit, where the carriers move in well ordered, nearly defect-free Cu-O planes. Comparable values of $`\mathrm{\Gamma }_D`$ can be obtained in doped semiconductors by the technique of ”modulation doping,” namely, by spatially separating the impurities, which provide the free carriers, from the planes where these latters are free to move. In BSCO the Cu-O planes seem then to be very clean, while the eventual impurities and defects are mostly distributed out-of-plane. The $`\omega ^1`$ dependence of $`\sigma (\omega )`$ in the mid infrared, often invoked to justify the need for an anomalous Drude model, is due to the tail of the peak centered at finite frequencies and should be ascribed to the bound charges. Similarly, the optical pseudogap (see Fig. 2) is a depression in $`\sigma (\omega )`$ created by the bound-charge absorption as this latter narrows and softens for decreasing $`T`$. This effect is required by the conservation of $`n_{eff}`$ with temperature, here fulfilled within a few percent. In this respect, any optical determination of the temperature $`T^{}`$ where the pseudogap starts opening seems questionable, as typically the FIR peak is observed even at $`T>T^{}`$. The present observations strongly support the coexistence of free and weakly bound charges in BSCO, responsible for the normal Drude absorption and the FIR peak, respectively. This latter is very similar to the polaronic feature observed in semiconducting cuprates at low doping. Its softening for decreasing $`T`$, followed by a saturation below $`T`$ 150 K, has already been attributed in Nd<sub>2-x</sub>Ce<sub>x</sub>CuO<sub>4</sub> to the formation of polaronic aggregates. This approach is confirmed by the present observations in BSCO, which are remarkably well fitted by a phase-separation model based on the scattering of the free carriers by dynamical arrays of weakly bound charges. We thank G. Balestrino for providing the sample here investigated. We are also indebted to C. Di Castro, R. Gonnelli, and M. Grilli for useful discussions.
warning/0001/math0001100.html
ar5iv
text
# Optimal Prediction of Stiff Oscillatory Mechanics ## 1 Stiff oscillatory mechanics There are many problems in classical mechanics where what can be computed is limited by the simultaneous presence of both fast and slow motion: some variables oscillate rapidly while others change slowly, so standard numerical methods can require a large number of time steps to give accurate answers. Stiffness of this type limits calculations of planetary motion, drift in high-frequency electronic oscillators, and the dynamics or large molecules . For instance, in molecular dynamics it is standard to model the motion of many atoms as a mechanical system with a Hamiltonian of the form $$H=\frac{1}{2}\underset{j=1}{\overset{N}{}}\frac{p_j^2}{2m_j}+V(q_1,\mathrm{},q_N)+\frac{1}{2}\underset{j=1}{\overset{N}{}}\underset{k=1}{\overset{N}{}}g_j(q)A_{jk}g_k(q)$$ (1) where $`(q_j,p_j)`$ are the coordinates and momenta of the atoms and $`N`$ is the number of atoms, commonly in the range $`10^4`$ to $`10^5`$. Here $`V`$ denotes a smoothly-varying potential energy of interaction among coordinates, the $`g`$’s are bond angles or interatomic spacings (functions of the coordinates), the $`m`$’s are masses, and $`A`$ is a matrix of spring constants. Such models are used to describe both the large-scale motion that takes place over milliseconds and also the rapid vibrational motions at chemical bonds which are measured in terahertz. In a recent paper, Stuart and Warren considered a particular stiff Hamiltonian problem of the form (1) that was originally meant to model a particle interacting with a heat bath , and they constructed numerical schemes that worked well with large time steps. They were able to compute the motion of slowly-varying quantities accurately, even when most of the dynamics was grossly underresolved in time (i.e., even when their time step was much longer than the periods of most normal modes of oscillation). This observation, that a scheme may be optimized to work well even when the resolution is poor, is similar to the results of optimal prediction ; optimal prediction is a method for reducing the resolution required to solve a large system of equations. A smaller system is constructed, designed to yield expectations of solutions of the larger system and to be computationally practical even when the larger system is not. Since Stuart and Warren have found schemes for some large, stiff systems that work with big time steps, it is natural to ask whether there are smaller systems of differential equations (just describing the slower modes) that would work at these big time steps. In this paper we show how optimal prediction may be applied to a class of large, stiff Hamiltonian systems like (1) to yield effective equations which are smaller and slower. We demonstrate the method on the Stuart-Warren model and on a generalization of it that more closely approximates realistic models of molecular dynamics. The benefits are longer time steps, lower dimensionality (hence fewer force evaluations per time step), and a systematic approach that may may be broadly applied. ## 2 Optimal prediction Optimal prediction is a method that takes a large system of differential equations together with a probability distribution for the dependent variables, and produces a smaller system of equations for the expectations of some selected variables while averaging over all the others. The method is described in . Error bounds for the method can be found in . Suppose we are given a large dynamical system $$\dot{u}_i=R_i(u_1,\mathrm{},u_N),i=1,\mathrm{},N$$ (2) for dependent variables $`u_1,\mathrm{},u_N`$, and we are also given a normalized probability density $`P(u_1,\mathrm{},u_N)`$ which is invariant under (2), $$\underset{j=1}{\overset{N}{}}\frac{P}{u_j}R_j(u_1,\mathrm{},u_N)0.$$ (3) The first step in the optimal prediction procedure is to identify “collective variables,” meaning a small number of functions of the dependent variables whose evolution we would like to predict. We denote these collective variables by $`v_1(u)\mathrm{}v_n(u)`$ where $`n<N`$. The idea in optimal prediction is to treat the $`u`$’s as random, treat their combinations in the $`v`$’s as known, and to estimate the rates of change of the $`v`$’s by conditional expectations. One writes out a formula for the rate of change of the $`v`$’s induced by (2), $$\dot{v}_\mu (u)=\underset{j=1}{\overset{N}{}}\frac{v_\mu }{u_j}R_j(u_1,\mathrm{},u_N)$$ (4) Then one uses $`P(u)`$ to compute the expectation of this expression subject to conditions that $`v_\mu (u)=\overline{v}_\mu `$ for some $`n`$ numbers $`\overline{v}_1\mathrm{}\overline{v}_n`$, $$\dot{v}_\mu _{\overline{v}_1\mathrm{}\overline{v}_n}=\frac{{\displaystyle \dot{v}_\mu (u)P(u)\underset{\nu =1}{\overset{n}{}}\delta (v_\nu (u)\overline{v}_\nu )du}}{{\displaystyle P(u)\underset{\nu =1}{\overset{n}{}}\delta (v_\nu (u)\overline{v}_\nu )du}}.$$ (5) Finally, one hypothesizes that the mean evolution of the $`v`$’s is approximated by the solutions $`\overline{v}_\mu (t)`$ of the new system, $$\dot{\overline{v}}_\mu (t)=\underset{i=1}{\overset{N}{}}\frac{v_\mu }{u_j}R_j(u_1,\mathrm{},u_N)_{\overline{v}_1(t)\mathrm{}\overline{v}_n(t)}$$ (6) The new system (6) is a closed system of equations for the $`\overline{v}`$’s, and it is $`n`$-dimensional instead of $`N`$-dimensional. Equation (6) approximates the evolution of the mean values of the $`v`$’s. The idea of the approximation is that at every moment in time, the $`u`$’s are distributed according to their invariant probability density subject to conditions on the values of collective variables. All that changes in time is the conditions, according to our hypothesis (6). Actually, if the $`v`$’s were given and the $`u`$’s were distributed according to a conditioned invariant distribution at time $`t=0`$, then at a future time $`t>0`$ the $`v`$’s would be indeterminate and the $`u`$’s would become distributed in some more general way. Average values of the $`v`$’s at all times $`t>0`$ would still be well-defined though, and they would be determined by the values of the $`v`$’s at $`t=0`$. The system (6) is meant to approximate such exact mean evolutions of collective variables from initial values. Although equation (6) is conjectural, some general results are known about its accuracy. First, it clearly gives an asymptotically exact prediction of mean futures for short times. Second, it appears in an exact formula for mean futures due to Zwanzig (recently studied by others ) which reveals corrections in terms of history integrals and noise-like functions which are statistically uncorrelated with the collective variables. Third, error bounds for the method have been established in the case of Hamiltonian dynamical systems . There are two technical challenges in the application of (6): collective variables must be selected, and the conditional expectations on the right-hand side must be explicitly evaluated, usually requiring approximations of the integrals in equation (5). Both steps are critical to accuracy. In complex problems, therefore, the best way to determine the usefulness of the approximation (6) is empirically: one generates large random ensembles of initial conditions for (2), integrates each initial condition, then averages the results to determine a mean future. One then compares the answer to an integral of (6). In the present paper, we will consider Hamiltonian equations where the dependent variables are canonical coordinate pairs $`(q_1,p_1)\mathrm{}(q_N,p_N)`$. Hamiltonian equations preserve the canonical probability density, $`e^H`$, so we will use this as our probability density. We assume that the first $`n`$ coordinate pairs $`(q_1,p_1)\mathrm{}(q_n,p_n)`$ are of interest, and we will take the remaining dynamical variables as random. The optimal prediction procedure is to take the full system of Hamilton’s equations, $$\dot{q}_j=\frac{H}{p_j},\dot{p}_j=\frac{H}{q_j},j=1,\mathrm{},N,$$ (7) discard the equations with indices $`j>n`$, and replace the right-hand sides of the remaining equations with their expectations with respect to $`e^H`$ conditioned by the selected variables: $$\dot{q}_\mu =\frac{H}{p_\mu }_n,\dot{p}_\mu =\frac{H}{q_\mu }_n,\mu =1,\mathrm{},n$$ (8) where $`_n`$ denotes the conditioned expectation, $$f_n=Z^1\underset{j=n+1}{\overset{N}{}}dq_jdp_je^Hf(q_1,\mathrm{},q_N;p_1,\mathrm{},p_N)$$ (9) with $`Z`$ a normalization constant. For any function $`f`$ of the canonical variables, $`f_n`$ is a function of $`q_1\mathrm{}q_n`$, $`p_1\mathrm{}p_n`$ only, so the $`2n`$-dimensional system of equations (8) is closed. The reduced system (8), the first approximation in optimal prediction, defines an approximate solution to a Liouville problem for the evolution of a probability measure on phase space. At least for short times, the system (8) is guaranteed to give the expectations of the selected variables, averaging over all possible initial data for the discarded variables. We need to evaluate the conditional expectations in (8). This is easy if $`e^H`$ is a Gaussian distribution (i.e., if $`H`$ is quadratic, or equivalently if the equations of motion are linear). If $`e^H`$ is not Gaussian, perturbative techniques are available to approximate its expectations by Gaussian expectations. Thus the following results for Gaussian distributions will be sufficient for our purposes, see for details. Let $`x_1,\mathrm{},x_N`$ be Gaussian random variables distributed with density $$P(x_1,\mathrm{},x_N)\mathrm{exp}\left(\frac{1}{2}\underset{j=1}{\overset{N}{}}\underset{k=1}{\overset{N}{}}x_jA_{jk}x_k+\underset{j=1}{\overset{N}{}}b_jx_j\right).$$ (10) We denote expectations with respect to this density by $``$, and $`x_i=_{j=1}^NA_{ij}^1b_j`$. Now suppose that $`x_1\mathrm{}x_n`$ are given for all $`n<N`$. The conditional expectations of $`x_{n+1}\mathrm{}x_N`$ conditioned by $`x_1\mathrm{}x_n`$ are denoted $`x_i_n`$, $`i=n+1,\mathrm{},N`$ and are given explicitly by $$x_i_n=x_i+\underset{\mu =1}{\overset{n}{}}\underset{\nu =1}{\overset{n}{}}A_{i\mu }^1M_{\mu \nu }^1(x_\nu x_\nu ),i=n+1,\mathrm{},N$$ (11) where $`M_{\mu \nu }=A_{\mu \nu }^1`$ for $`\mu ,\nu =1,\mathrm{},n`$ and $`M^1`$ is the inverse of the $`n\times n`$ (not $`N\times N`$) matrix $`M`$. The conditioned covariances, $`\text{Cov}_n(x_i,x_j)=x_ix_j_nx_i_nx_j_n`$ are given in terms of the unconditioned expectations $`\text{Cov}(x_i,x_j)=x_ix_jx_ix_j`$ by $$\text{Cov}_n(x_i,x_j)=\text{Cov}(x_i,x_j)\underset{\mu =1}{\overset{n}{}}\underset{\nu =1}{\overset{n}{}}A_{i\mu }^1M_{\mu \nu }^1A_{\nu j}^1.$$ (12) The conditioned expectation of any polynomial in $`x_1\mathrm{}x_N`$ may be found from these formulae by Wick’s theorem. ## 3 Generalizations of the Stuart-Warren experiments Stuart and Warren (see also , and ) considered a one-dimensional collection of particles connected by springs. There was one distinguished particle with mass $`1`$, coordinate $`Q`$ and momentum $`P`$. The distinguished particle was connected by springs of spring constant $`k`$ to $`N`$ other particles with masses $`k/j^2`$, coordinates $`q_j`$ and momenta $`p_j`$, $`j=1\mathrm{}N`$, representing a heat bath. The motion of this collection of particles and springs is defined by the Hamiltonian $$\begin{array}{c}H(Q,P;q_1,\mathrm{},q_N;p_1,\mathrm{},p_N)\hfill \\ \hfill =\frac{1}{2}(V(Q)+P^2)+\underset{j=1}{\overset{N}{}}\left[\frac{p_j^2}{2m_j}+\frac{1}{2}k(Qq_j)^2\right]\end{array}$$ (13) where $`(Q,P)`$ and $`(q_j,p_j)`$ are canonically conjugate dynamical variables for $`j=1,\mathrm{},N`$ and $`m_j=k/j^2`$. The equations of motion are $`\dot{Q}`$ $`=P`$ $`\dot{P}`$ $`=V^{}(Q)+k{\displaystyle \underset{j=1}{\overset{N}{}}}(q_jQ)`$ (14) $`\dot{q}_j`$ $`=p_j/m_j`$ $`\dot{p}_j`$ $`=k(Qq_j),j=1,\mathrm{},N`$ This system is of the form (1) (with an extra pair of coordinates $`(Q,P)`$), and it is chosen so that fast and slow motion are separated: lighter particles will move faster, heavier particles will move slower, and the mass $`m_j`$ goes down as $`j`$ goes up. A central result of is that if all the heat bath particles start out randomly, with statistics determined by the canonical distribution, then in the limit $`N\mathrm{}`$ the coordinate of the distinguished particle obeys the stochastic equation, $$\ddot{Q}+\frac{k\pi }{2}\dot{Q}+V^{}(Q)\frac{k}{2}Q=F$$ (15) where $`F(t)`$ is a stochastic process related to white noise. This equation for $`Q`$ is remarkable because it makes no reference to the history of $`Q`$—it is a differential equation, not an integro-differential equation. In a general Hamiltonian problem, if one variable $`Q`$ is fixed initially and the others are random, at future times there is no time-invariant relationship among the expectation of $`Q`$ and its time derivatives . The first approximation of optimal prediction (8) may be characterized as the assumption that the values of the selected variables do determine their own future expectations. In general this assumption is not exactly true, but in the Stuart-Warren model it is true exactly in the $`N\mathrm{}`$ limit. Stuart and Warren proceeded to integrate their model with large time steps. If $`Q`$ were fixed, then each $`q_j`$ would oscillate harmonically with frequency $`\omega _j=j`$. This implies that a discretization of the $`2N+2`$ equations (14) would be resolved in time if $`N\mathrm{\Delta }t1`$. If this condition on $`\mathrm{\Delta }t`$ were violated, then the result of the computation would depend on how the equations were discretized. The intriguing result of is that some schemes will give the right evolution for $`Q`$ and $`P`$ when $`N\mathrm{\Delta }t1`$ and others will not. For instance, if the scheme is $`{\displaystyle \frac{Q^{n+1}Q^n}{\mathrm{\Delta }t}}`$ $`=P^{n+1}`$ $`{\displaystyle \frac{P^{n+1}P^n}{\mathrm{\Delta }t}}`$ $`=V(Q^n)+k{\displaystyle \underset{j=1}{\overset{N}{}}}(q_j^{n+\sigma }Q^n)`$ (16) $`{\displaystyle \frac{q_j^{n+1}q_j^n}{\mathrm{\Delta }t}}`$ $`=p_j^{n+1}/m_j`$ $`{\displaystyle \frac{p_j^{n+1}p_j^n}{\mathrm{\Delta }t}}`$ $`=k(Q^nq_j^n)j=1,\mathrm{},N`$ then $`\sigma =0`$ (a symplectic method) gives the right answer for $`Q`$ and $`P`$, but $`\sigma =1`$ (another convergent method) does not. For concreteness, we pick $`V(Q)=\frac{1}{2}Q^2`$. Since $`H`$ in (13) is then quadratic, the canonical probability density is Gaussian, and formula (11) gives the conditioned expectations as $$q_j_n=Q,p_j_n=0(n<jN).$$ (17) Taking the conditional expectations of the right-hand sides of (14) and evaluating them using these results, we find that the equations of optimal prediction are $`\dot{Q}`$ $`=P`$ $`\dot{P}`$ $`=Q+k{\displaystyle \underset{\mu =1}{\overset{n}{}}}(q_\mu Q)`$ (18) $`\dot{q}_\mu `$ $`=p_\mu /m_\mu `$ $`\dot{p}_\mu `$ $`=k(Qq_\mu ),\mu =1,\mathrm{},n`$ These are identical in form to the original equations (14). It comes as no surprise, therefore, that the motion of $`Q`$ can be computed with large $`\mathrm{\Delta }t`$: pick the $`\mathrm{\Delta }t`$ desired, find an $`nN`$ such that $`n\mathrm{\Delta }t1`$, and perform a resolved integration of (18) with this $`n`$ and $`\mathrm{\Delta }t`$. Reasonable approximations for the selected variables are guaranteed, at least for short times. Figure 1 shows a fully-resolved calculation ($`N\mathrm{\Delta }t=10^2`$) of $`P(t)`$ starting from $`P(0)=0`$, $`Q(0)=1.5`$, with $`q_j(0)`$ and $`p_j(0)`$ chosen randomly from the canonical ensemble (i.e., chosen with probability density $`e^H`$) conditioned by $`Q(0)`$ and $`P(0)`$. It also shows the solution to the same problem as computed by a resolved integration of (18), which was achieved with $`n\mathrm{\Delta }t=10^2`$. The optimal prediction calculation accurately duplicates the low-frequency behavior of the exact solution, and it does so in fewer dimensions with a larger time step. In this case, with $`N=10^4`$ and $`n=10^2`$, the optimal prediction curve was about $`10,000`$ times faster to compute than the resolved solution. The optimal prediction has the further advantage that it did not use the initial data $`q_{n+1}(0)\mathrm{}q_N(0)`$, $`p_{n+1}(0)\mathrm{}p_N(0)`$ and may claim to be an average answer over all possible values of these data. ## 4 More general models Realistic applications, such as molecular dynamics, involve more complex interactions than are present in the model (14). In particular, we may expect that every particle would interact with every other, and that the interactions would be nonlinear. We therefore consider a generalization of the model (14) where every $`q_1\mathrm{}q_N`$ is coupled to every other $`q_1\mathrm{}q_N`$ by a spring, and the springs are nonlinear: $$\begin{array}{c}H(q_1,\mathrm{},q_N;p_1,\mathrm{},p_N)\hfill \\ \hfill =\underset{j=1}{\overset{N}{}}\frac{p_j^2}{2m_j}+\frac{1}{2}k^{(2)}\underset{j=1}{\overset{N}{}}\underset{l=j+1}{\overset{N}{}}(q_jq_l)^2+\frac{1}{4}k^{(4)}\underset{j=1}{\overset{N}{}}\underset{l=j+1}{\overset{N}{}}(q_jq_l)^4\end{array}$$ (19) $$\begin{array}{cc}\hfill \dot{q}_j& =p_j/m_j\hfill \\ \hfill \dot{p}_j& =k^{(2)}\underset{l=1}{\overset{N}{}}(q_jq_l)k^{(4)}\underset{l=1}{\overset{N}{}}(q_jq_l)^3\hfill \end{array}\}j=1,\mathrm{},N.$$ (20) This model makes no reference to a distinguished particle; each one of the $`N`$ particles interacts with all of the others through the same potential energy, which is parameterized by the new spring constants $`k^{(2)}`$ and $`k^{(4)}`$. We derive the optimal prediction equations of the system (20) for $`q_1\mathrm{}q_n`$, $`p_1\mathrm{}p_n`$ by averaging over $`q_{n+1}\mathrm{}q_N`$, $`p_{n+1}\mathrm{}p_N`$. Since the interactions are now nonlinear, the probability density $`e^H`$ is no longer Gaussian, so we must work harder to evaluate the conditioned expectations. Hald has observed, as reported in , that optimal prediction equations of the form (8) are always Hamiltonian, and that their Hamiltonian is $$H^{}(q_1,\mathrm{},q_n;p_1,\mathrm{},p_n)=\mathrm{log}\left(\underset{j=n+1}{\overset{N}{}}dq_jdp_je^H\right).$$ (21) We may therefore approximate the conditioned expectations of (8) by first approximating $`H^{}`$, and then deriving (8) by differentiation: $$\dot{q}_\mu =\frac{H^{}}{p_\mu },\dot{p}_\mu =\frac{H^{}}{q_\mu },\mu =1,\mathrm{},n.$$ (22) We decompose $`H`$ into its quadratic part plus its higher-order part, $`H`$ $`=H_0+H_1`$ (23) $`H_0`$ $`={\displaystyle \underset{j=1}{\overset{N}{}}}{\displaystyle \frac{p_j}{2m_j}}+{\displaystyle \frac{k^{(2)}}{2}}{\displaystyle \underset{j=1}{\overset{N}{}}}{\displaystyle \underset{l=j+1}{\overset{N}{}}}(q_jq_l)^2`$ $`H_1`$ $`={\displaystyle \frac{k^{(4)}}{4}}{\displaystyle \underset{j=1}{\overset{N}{}}}{\displaystyle \underset{l=j+1}{\overset{N}{}}}(q_jq_l)^4`$ and proceed by determining $`H^{}`$ perturbatively as a power series in $`k^{(4)}`$. An alternate method for perturbative treatment of optimal prediction is described in . Hald’s formula (21) implies $`H^{}`$ $`=\mathrm{log}\left({\displaystyle \underset{j=n+1}{\overset{N}{}}dq_jdp_je^{H_0}}\right)\mathrm{log}\left({\displaystyle \frac{_{j=n+1}^ndq_jdp_je^{H_0}e^{H_1}}{_{j=n+1}^Ndq_jdp_je^{H_0}}}\right)`$ (24) $`=\text{(}H_0\text{-part)}\mathrm{log}e^{H_1}_{n,0}`$ where the new average, $`_{n,0}`$ denotes an average with respect to the conditioned Gaussian measure, defined just as in the definition (9) but with $`H_0`$ replacing $`H`$. The “($`H_0`$-part)” term would be the effective Hamiltonian if $`H_1`$ were zero, and it contributes linear terms to the equations of motion which are easily evaluated by the regression formula (11). The other term in (24) is equal to a power series in $`k^{(4)}`$, $$\mathrm{log}e^{H_1}_{n,0}=\underset{m=1}{\overset{\mathrm{}}{}}\frac{(1)^m}{m!}H_1^m_{n,0}^{(c)}$$ (25) where $`H_1^m_{n,0}^{(c)}`$ denotes the $`m`$-th cumulant of $`H_1`$ with respect to the conditioned Gaussian measure. Each cumulant in this series may be evaluated by Wick’s theorem, where only “connected” pairings (in the sense of perturbation theory in physics) are included. To first order in $`k^{(4)}`$, we need to evaluate $`H_1^1_{n,0}^{(c)}`$ $`=H_1_{n,0}`$ (26) $`={\displaystyle \frac{k^{(4)}}{4}}{\displaystyle \underset{j=1}{\overset{N}{}}}{\displaystyle \underset{l=j+1}{\overset{N}{}}}(q_jq_l)^4_{n,0}`$ $`={\displaystyle \frac{k^{(4)}}{4}}\left[{\displaystyle \underset{\mu =1}{\overset{n}{}}}{\displaystyle \underset{\nu =\mu +1}{\overset{n}{}}}(q_\mu q_\nu )^4+{\displaystyle \underset{\mu =1}{\overset{n}{}}}{\displaystyle \underset{l=\mu +1}{\overset{n}{}}}(q_\mu q_l)^4_{n,0}\right]+\text{(constant)}`$ where “(constant)” denotes terms that are independent of $`q_1,\mathrm{},q_n`$ and $`p_1,\mathrm{},p_n`$ (and therefore do not affect equations of motion). The average $`_{n,0}`$ may be deduced from the expectations, $$\begin{array}{cc}\hfill q_j_{n,0}& =\frac{1}{n}\underset{\mu =1}{\overset{n}{}}q_\mu \hfill \\ \hfill \text{Cov}_0(q_j,q_l)& =\frac{1}{Nk^{(2)}}(1+\delta _{jl})\hfill \end{array}j,l=n+1,\mathrm{},N$$ (27) together with Wick’s theorem. The result for $`H^{}`$, to first order in $`k^{(4)}`$, is $`H^{}={\displaystyle \underset{\mu =1}{\overset{n}{}}}{\displaystyle \frac{p_\mu ^2}{2m_\mu }}`$ $`+{\displaystyle \frac{C_2}{2}}{\displaystyle \underset{\mu =1}{\overset{n}{}}}{\displaystyle \underset{\mu =\nu +1}{\overset{n}{}}}(q_\mu q_\nu )^2`$ (28) $`+{\displaystyle \frac{C_4}{4}}{\displaystyle \underset{\mu =1}{\overset{n}{}}}{\displaystyle \underset{\mu =\nu +1}{\overset{n}{}}}(q_\mu q_\nu )^4`$ $`+{\displaystyle \frac{D_4}{4}}{\displaystyle \underset{\mu =1}{\overset{n}{}}}\left(q_\mu {\displaystyle \frac{1}{n}}{\displaystyle \underset{\nu =1}{\overset{n}{}}}q_\nu \right)^4+O\left(k^{(4)}\right)^2`$ where the coupling constants to this order in $`k^{(4)}`$ are $$\begin{array}{cc}\hfill C_2& =\frac{N}{n}k^{(2)}+3\frac{(Nn)(n+1)}{Nn}\frac{k^{(4)}}{k^{(2)}}\hfill \\ \hfill C_4& =k^{(4)}\hfill \\ \hfill D_4& =k^{(4)}(Nn)\hfill \end{array}.$$ (29) We differentiate (28) to obtain the optimal prediction equations for the new system (20) to $`O(k^{(4)})^2`$, $$\begin{array}{cc}\hfill \dot{q}_\mu & =p_\mu /m_\mu \hfill \\ \hfill \dot{p}_\mu & =C_2\underset{\nu =1}{\overset{n}{}}(q_\mu q_\nu )C_4\underset{\nu =1}{\overset{n}{}}(q_\mu q_\nu )^3\hfill \\ & D_4\frac{1}{n}\underset{\nu =1}{\overset{n}{}}\left[\left(q_\mu \frac{1}{n}\underset{\sigma =1}{\overset{n}{}}q_\sigma \right)^3\left(q_\nu \frac{1}{n}\underset{\sigma =1}{\overset{n}{}}q_\sigma \right)^3\right]\hfill \end{array}\mu =1,\mathrm{},n.$$ (30) We performed a more rigorous test of this new model, comparing it to an actual mean evolution. The results are shown in Figure 2. We once again picked $`q_1\mathrm{}q_n`$, $`p_1\mathrm{}p_n`$ ($`n=10`$) from the canonical distribution $`e^H`$ for $`N`$ particles ($`N=1000`$ at $`k^{(2)}=1`$ and $`k^{(4)}=0.1`$). We then generated an ensemble of $`100`$ sets of values for $`q_{n+1}\mathrm{}q_N`$, $`p_{n+1}\mathrm{}p_N`$ from the canonical distribution conditioned by $`q_1\mathrm{}q_n`$, $`p_1\mathrm{}p_n`$, and for each set integrated the equations (20). Averaging over all $`100`$ solutions yielded the solid curve for $`p_1(t)`$. We then discarded the ensemble and used the original $`q_1\mathrm{}q_n`$, $`p_1\mathrm{}p_n`$ as initial conditions for the reduced system (30), which we integrated with $`\mathrm{\Delta }t=10^2/n=1/N`$. This $`\mathrm{\Delta }t`$ is small enough to resolve the reduced dynamics but much too large to resolve the original dynamics. The solution for $`p_1(t)`$ from (30) is the dashed curve. Finally, for comparison we performed the naive experiment of simply truncating the big system (20) to $`n`$ degrees of freedom, effectively ignoring the lighter particles without changing the interactions. This produced the dot-dashed curve. The figure shows that the reduced system accurately predicts the average evolution of $`p_1(t)`$, and it does so with $`1`$ percent of the degrees of freedom and time steps that are $`100`$ times larger. The naive experiment shows that the new couplings are critical to the answer. Since forces must be evaluated $`N(N1)/2`$ times per time step for $`N`$ particles, optimal prediction speeds up the calculation of $`p_1(t)`$ in this case by about a factor of about $`10^6`$. ## 5 Conclusions We have shown that optimal prediction may be applied to large, stiff Hamiltonian systems of differential equations to make new systems that are smaller, better-conditioned, and approximate the original equations in the mean. We have demonstrated that the method gives accurate answers while allowing larger time steps and requiring fewer force evaluations. ## 6 Acknowledgements The author thanks Profs. A. Chorin, O. Hald, R. Kupferman, and A. Stuart for helpful discussions.
warning/0001/cond-mat0001192.html
ar5iv
text
# Theory of Chiral Imprinting ## Abstract We present a continuum model for a nematic elastomer network formed in a chiral environment, for instance in the presence of a chiral solvent. When this environment is removed, the network can retain some memory of its chiral genesis. We predict the residual chiral order parameter for a number of possible scenarios, and go on to examine the robustness (stability) of the imprinted chirality. We show that a twist-untwist transition can take place, which determines whether the imprinting has been successful. A transition is via a coarsening of the helical director pattern and a lengthening of its pitch. Finally, the effect due to a subsequent swelling by an achiral solvent, or by a solvent of differing chirality, is considered. PACS numbers: 61.30.-v, 61.41.+e, 78.20.Ek Nematic elastomers combine the properties of a liquid crystal and those of a conventional rubber. This synergy gives rise to novel material behaviour, which in turn has stimulated much research in past years . From a general symmetry argument, de Gennes first suggested that chirality may be introduced to such an elastomer by simply forming it in a chiral solvent. The originally achiral liquid crystalline polymer would then remember the induced chirality after crosslinking, even when the solvent is removed or replaced with an achiral one. This is the case of chiral imprinting, which potentially can open up an entirely new way of producing materials of specified optical properties. Chiral imprinting, in principle, is akin to crosslinking a nematic polymer under an external magnetic or mechanical field where the monodomain state is also permanently imprinted. Experimentally chiral imprinting was studied long ago as a function of solvent exchange. Recently imprinting has been studied as a function of both solvent removal and temperature. Another measure of imprinting occurs in intrinsically cholesteric networks. On temperature changes that would cause a substantial pitch variation in a non-crosslinked cholesteric polymer melt, the corresponding network suffers essentially no variation - see for example Fig. 8 of reference , where also many other references to cholesteric elastomers are given. In this paper, we analyse chiral imprinting, and predict the retained chiral properties of the elastomer when the initial chiral environment of crosslinking is altered. Gradients of director variation are modeled within continuum Frank nematic elasticity. The nematic elastomer penalty for rotation of the director relative to the solid matrix is described in a fully non-linear, rubber-elastic manner since rotation can be large. A nematic liquid crystal has a mobile director $`𝐧`$, the gradient of which incurs a Frank energy . For the free energy density of a cholesteric liquid crystal, the twist term is modified by a pitch wave number, $`q_0`$: $$f_c=\frac{1}{2}\left[K_1(\mathbf{}𝐧)^2+K_2(𝐧(\mathbf{}\times 𝐧)+q_0)^2+K_3((𝐧\mathbf{})𝐧)^2\right]$$ where $`K_{1,2,3}`$ respectively measure the energy penalty for splay, twist and bend, the three possible modes of the nematic director distortion. In a pure twisting, cholesteric conformation of the director, we can drop the terms involving $`K_{1,3}`$. These other contributions arise if the director moves away from the helical plane and tilts toward the pitch axis. The Frank energy can be viewed as a continuum description with higher order spatial derivatives truncated, and it suffices for cases where director varies slowly over a nematic coherence length ($`10`$nm). When liquid crystalline polymers are crosslinked into a rubber network, additional constraints on $`𝐧`$ arise in the form of director anchoring to the network. Anchoring manifests itself with an extra energy cost for a uniform director rotation $`𝝎`$ relative to a local rotation of the elastic matrix $`𝛀`$. The energy density is $`\frac{1}{2}D_1[(𝛀𝝎)\times 𝐧]^2`$. A second term couples the relative director-matrix rotation to the shearing part of the elastic strain, $`𝝀`$, that is $`D_2𝐧.𝝀.[(𝛀𝝎)\times 𝐧]`$. For such strains to rotate the director in a cholesteric, they would have to vary along the helical axis. By the requirement of elastic compatibility this introduces secondary shears. One can show that these are prohibitively expensive for this geometry. We accordingly ignore $`𝝀`$ terms. Consider an initially regular cholesteric helix along the $`z`$ axis, Fig. 1. $`\theta `$ is the azimuthal angle the local director makes within the $`xy`$ plane and is initially $$\theta _0(z)=q_0z$$ (1) where $`q_0`$ is the chiral pitch wavenumber. One can generalise the $`D_1`$ term to the large angle limit by considering a molecular nematic rubber elastic model . It is no longer $`\frac{1}{2}D_1(\theta \theta _0)^2`$ but $`\frac{1}{2}D_1\mathrm{sin}^2(\theta q_0z)`$ which has the correct locally nematic symmetry of $`𝐧𝐧`$. The energy for an elastomer formed under a cholesteric solvent which is subsequently replaced with an achiral one is then: $$F=𝑑z\frac{1}{2}[K_2\theta ^2+D_1\mathrm{sin}^2(\theta q_0z)]$$ (2) where indicates $`d/dz`$. The first term in $`F`$ wishes to remove the twist in $`𝐧`$ since now the chiral imperative of the solvent is removed the usual Frank twist penalty is fully incurred. However the second term insists that $`𝐧`$ is anchored to a helix thanks to the crosslinks being formed under cholesteric conditions. The initial network polymers are taken to be achiral, thus $`q_0`$ is only induced and can be tuned with the choice of the chiral solvent we subject our elastomer to at formation. If achiral solvent is used when crosslinking, $`q_0`$ could simply be zero, and we retrieve the description of more conventional nematic elastomers . The two limits of the energy density are (i) the perfectly twisted cholesteric state $`f=\frac{1}{2}K_2q_0^2`$ where the current pitch wavevector is unchanged from $`q_0`$, and (ii) the untwisted state $`f=\frac{1}{4}D_1`$, the additional factor of $`\frac{1}{2}`$ arising from the averaging of $`\mathrm{sin}^2`$ over one period. Thus, crudely, we expect the director to be twisted if $$K_2q_0^2<D_1/2$$ (3) and untwisted otherwise, this balance being tuneable since $`K_2`$, $`D_1`$ and $`q_0^2`$ vary relatively to each other with temperature, degree of crosslinking and swelling and the presence of additional chiral agents. Eq. (3) anticipates the physics of our detailed results: highly twisted states ($`q_0`$ large) or systems with a large twist constant $`K_2`$ will pay a very high Frank penalty on loss of spontaneous twist arising from the loss of the chiral solvent. A large combination $`K_2q_0^2`$ will overcome the anchoring $`D_1`$ and the elastomer will untwist - imprinting will be lost. Weakly twisted elastomers with weak twist constants will not overcome director anchoring and imprinting will remain. We now analyze Eq. (2) for details of the phase behaviour. To simplify matters, we begin with the following substitutions: $$\varphi =q_0z\theta +\pi /2;u=z/\xi ;\xi =\sqrt{K_2/D_1};\alpha =\xi q_0$$ the angle $`\varphi `$ describes the variation away from (or modulation of) the original helical pattern, that is we are now in a rotating frame of reference. Lengths are reduced by the nematic rubber penetration depth $`\xi `$, the natural length scale in the problem. The parameter $`\alpha `$ is a non-dimensional measure of the nematic length relative to the chiral pitch, and the condition (3) is equivalent to $`\alpha 1/\sqrt{2}`$. Following the remarks below Eq. (3), we expect imprinting to be lost at large $`\alpha `$ and retained at small $`\alpha `$. Dropping a constant arising from the variable change $`\theta \varphi `$, the reduced energy (per unit cross-section area perpendicular to the pitch axis) is: $$\stackrel{~}{F}=\frac{2F}{D_1\xi }=𝑑u[(\dot{\varphi }\alpha )^2\mathrm{sin}^2\varphi ]$$ (4) where $`(\dot{})`$ signifies $`d/du`$. If we make the analogy to Lagrangian dynamics, the integrand is the Lagrangian density ($`L=TV`$) of a particle in a potential given by $`\mathrm{sin}^2\varphi `$. The problem also now resembles the problem of an electric or magnetic field applied perpendicularly to the helix of a liquid cholesteric, solved long ago by de Gennes and Meyer . In reduced terms, the $`\mathrm{sin}^2\varphi `$ is like the field competing with the natural chirality $`\alpha `$. In terms of the original problem, Eq. (2), it is as if a naturally untwisted nematic has a spatially chiral electric field “$`D_1\mathrm{"}`$ applied to it in an attempt to induce a twist. The first integral of the Euler-Lagrange equation corresponding to Eq. (4) , the elliptic equation or often called in the literature the Sine-Gordon equation, leads to: $$\dot{\varphi }^2+\mathrm{sin}^2\varphi =c^2$$ (5) where $`c^2`$ is the integration constant, which has the physical interpretation of total energy. The analogy is helpful, and before we examine the details we already foresee two scenarios: If $`c^2<1`$, the particle does not have sufficient energy to overcome the barriers in $`\mathrm{sin}^2\varphi `$ potential and is therefore localized, i.e. $`\pi /2<\varphi <\pi /2`$. This corresponds to the case where cholesteric pitch is largely maintained, with $`\varphi (u)`$ only introducing small modulations to the director orientation. If $`c^2>1`$, our particle can climb out of the potential valleys and travel freely; this corresponds to the case of winding/unwinding the cholesteric twists in director orientation. Below, we determine $`c`$ in terms of $`\alpha `$. The localised limit, $`c^2<1`$. Our particle oscillates between two values of $`\varphi _m=\pm \mathrm{arcsin}c`$. We accordingly introduce a new variable $`\beta `$ in the interval $`\beta [\frac{\pi }{2},\frac{\pi }{2}]`$, so that $`\mathrm{sin}\varphi c\mathrm{sin}\beta `$. Rewriting the derivative $`\dot{\varphi }`$ in terms of $`\dot{\beta }`$ and returning it to Eq. (5) reduces this equation to one of the standard elliptic form: $`\dot{\beta }=\sqrt{1c^2\mathrm{sin}^2\beta }`$. The period of the oscillatory motion is found to be: $$T_1=2_0^{\frac{\pi }{2}}\frac{d\beta }{\dot{\beta }}=2𝒦(c).$$ (6) Here $`𝒦(c)`$, and later $`(c)`$, are the complete elliptic integrals of the first and second kind respectively. The period $`T_1`$ gives, in units of the characteristic length scale $`\xi `$, the spatial repeat distance of our $`\varphi `$-modulation of the original cholesteric angle $`q_0z`$. The reduced energy of a period can be obtained from Eq. (4) as: $`\stackrel{~}{F}_1(T_1)`$ $`=`$ $`2[(c^2+\alpha ^22)𝒦(c)+2(c)].`$ (7) In order to compared the stability of various states, we require the reduced energy density. In the localized regime, $`c^2<1`$, we denote this by $`g_1(c)`$: $$g_1=\stackrel{~}{F}_1(T_1)/T_1=c^2+\alpha ^22+\frac{2(c)}{𝒦(c)}.$$ (8) We now require the energy density in the traveling regime. The traveling limit $`c^2>1`$. The modulation period down the original helix corresponds to the time taken for our particle to travel from one peak to the next in the potential. It is calculated using Eq. (5): $$T_2=2_0^{\frac{\pi }{2}}\frac{d\varphi }{\dot{\varphi }}=2k𝒦(k)$$ (9) where $`k=1/c`$, and thus $`k<1`$. The reduced energy of a period is: $`\stackrel{~}{F}_2(T_2)`$ $`=`$ $`2[(\alpha ^2c^2)k𝒦(k)+2c(k)\alpha \pi ].`$ (10) The corresponding energy density, $`g_2(c)`$ for $`c^2>1`$, is: $$g_2=\stackrel{~}{F}_2(T_2)/T_2=\alpha ^2c^2+\frac{2}{k𝒦(k)}\left[\frac{(k)}{k}\frac{\alpha \pi }{2}\right].$$ (11) Combining $`g_1(c)`$ and $`g_2(c)`$, Eq.s (8, 11), we now have the energy density $`g(c)`$ for the entire range of $`c`$. Fig. 1 illustrates the energy density plots for three different values of of the chiral strength, described by the parameter $`\alpha `$. We can see that at the transition point $`c=1`$ the two energy densities approach each other. However, the densities are not smoothly joined, forming a cusp which turns from pointing downward to upward upon the parameter $`\alpha `$ increasing past a critical value $`\alpha _c=2/\pi `$. Invoking the identities for differentials of complete elliptic integrals: $$\frac{d𝒦}{dk}=\frac{}{k(1k^2)}\frac{𝒦}{k};\frac{d}{dk}=\frac{𝒦}{k}$$ we can minimise the reduced free energy with respect to $`k`$, or equivalently $`c`$, by setting $`dg/dc=0`$. As evident from Fig. 2, minima only exist in the region $`c^2>1`$, that is in $`g_2(c)`$, and only when $`\alpha >\alpha _c`$. There are no minima in $`g_1`$. The analogous problem of a cholesteric liquid in the presence of an electric field is similar. The condition $`dg/dc=0`$ fixes $`c`$ since, for a given $`\alpha `$, $`k`$ or $`c`$ must satisfy: $$\alpha =\frac{2(k)}{\pi k}\mathrm{for}\alpha >\alpha _c=2/\pi .$$ (12) The period of $`\varphi (u)`$ modulation is $`T=2k𝒦(k)`$ in units of $`\xi `$. For each such period along the helical pitch axis, $`\varphi `$ increases by $`\pi `$, the director unwinds once, and therefore one loses $`1`$ of the $`T\xi q_0/\pi 2\alpha k𝒦(k)/\pi `$ twists imprinted over the interval $`T`$. The imprinting efficiency is given by the fractional number of twists lost: $$e_0=\frac{2\alpha k𝒦(k)\pi }{2\alpha k𝒦(k)}.$$ (13) Provided that $`e_0`$ is close to unity, much of the chirality can be preserved. For small chiral power $`\alpha <\alpha _c=2/\pi `$, the minimising condition $`dg/dc=0`$ has no solution and the minimum free energy occurs exactly at $`c=1`$. The cusp has the energy density $`g=(2/\pi )^21`$, and a logarithmically divergent period $`T`$ which implies $`e_0=1`$. The director gets arrested, attempting to untwist but never actually manages a full turn within a finite distance. The imprinting is therefore successful. For chiral power $`\alpha >\alpha _c=2/\pi `$, minima are found for $`c>1`$. The period is no longer infinite and twists are lost. For large $`\alpha `$, i.e. the nematic penetration depth large compared to the cholesteric pitch, we can expand the elliptic function for small $`k`$ to find $`\alpha c(1k^2/43k^4/128\mathrm{})`$. Alternatively $`k1/\alpha (11/4\alpha ^2\mathrm{})`$ and we find a period of $`T=2k𝒦(k)\pi /\alpha `$ in units of $`\xi `$. Thus for every actual distance of $`\pi \xi /\alpha =\pi /q_0`$, $`\varphi `$ accumulates an increment of $`\pi `$. That is, $`e_00`$, corresponding eventually to the case of complete unwinding of the imprinted helical pattern. In writing equation (2), we have assumed that the chiral solvent was completely replaced with an achiral one as was the case in experiment . However, we can trivially generalise $`F`$ to the case where the current solvent is chiral: $$F=𝑑z\frac{1}{2}[K_2(\theta ^{}q)^2+D_1\mathrm{sin}^2(\theta q_0z)],$$ (14) where $`q`$ is the wave number that the solvent would introduce in the absence of crosslinks. Our previous case corresponds to $`q=0`$. The mathematical procedure remains applicable, with an adjustment of the parameter $`\alpha =\xi q_0`$ to: $$\alpha ^{}=\xi (q_0q)$$ which indicates a chiral environment of the same handedness can help preserve the chirality by reducing the value of $`\alpha `$ (negative values of $`\alpha `$ are mathematically equivalent to $`\alpha `$, if we switch the sign of the $`\varphi `$ modulation). A chiral environment of the other handedness, $`q<0`$, instead increases $`\alpha `$ and has the opposite effect. It reduces imprinting, as one might expect. By simply varying $`q`$, via the solvent composition, we have a powerful way to map out the wind-unwinding transition. Another interesting case is the induction of a cholesteric state in an initially uniform nematic network upon introducing a chiral solvent . That is $`q_0=0`$ and $`\alpha ^{}=\xi q`$, where $`q`$ is the chiral pitch wave number due to the chiral solvent. In this case, $`e_0`$ given in Fig. 2 can be interpreted as the network resistance to imprinting - until the point $`\alpha ^{}`$ reaches $`2/\pi `$ there will be no helicity induced in the network at all, and thereafter it rises with $`\alpha ^{}`$. Finally, by varying the amount of solvent in the crossed network, we can also tune, i.e. contract or expand, the volume relative to its original: $`V\beta V`$. With this, we expect from molecular interpretations, $`K_2K_2`$ for a special case where the local nematic order is preserved (the nematic order variation due to the solvent or temperature will be examined elsewhere), but $`D_1D_1/\beta `$ as it depends on crosslink density, and finally assuming an isotropic expansion/contraction, $`q_0q_0\beta ^{1/3}`$. Thus we have $$\alpha \alpha \beta ^{1/6},$$ so swelling can, albeit weakly, increase the parameter $`\alpha `$ in our model and discourage the preservation of the imprinted chirality. In conclusion, we have proposed a continuum model for chiral imprinting in nematic elastomers. The model predicts the residual chirality when the chiral solvent is removed. A twist-untwist transition emerges from the theory, and an experimental verification of this transition would be a useful test of the theory presented here. Further work is focussed on the mechanical properties of networks; we expect an external mechanical field will have substantial influence on the imprinting of chirality. We thank E. M. Terentjev, R. B. Meyer, and M. E. Cates for useful discussions. The problem was first suggested to us by H. Finkelmann. YM is grateful to St John’s College, Cambridge for a research fellowship.
warning/0001/hep-th0001076.html
ar5iv
text
# Generalized two-dimensional Yang-Mills theory is a matrix string theory ## 1 INTRODUCTION In the early seventies ’t Hooft established that in the limit of a large number $`N`$ of colours gauge theories are dominated by planar diagrams, namely by Feynman diagrams with the topology of a sphere, and that diagrams of genus $`g`$ are suppressed by a factor $`[\frac{1}{N}]^{2g2}`$. The expansion in powers of $`\frac{1}{N}`$ can thus be reinterpreted as a perturbative string expansion, with string coupling constant $`\frac{1}{N}`$. In two space-time dimensions, gauge fields have no physical degrees of freedom and pure Yang-Mills theory (YM2) is invariant under area preserving diffeomorphisms, thus revealing its almost topological nature. As a consequence of these features it can be exactly solved on arbitrary Riemann surfaces, both in a lattice and continuum formulation. The interpretation in terms of a string theory of this solution was provided in a series of beautiful papers where it was shown that the coefficients of the $`\frac{1}{N}`$ expansion of the YM2 partition function count the number of maps (without foldings) from a two-dimensional world-sheet of genus-$`g`$ to the Euclidean space-time, which is a Riemann surface of genus $`G`$. In two dimensions the number of maps is the number of string configurations and, as each map is weighed by the area of its world sheet, one can conclude that in this limit YM2 is a two-dimensional string theory. A string theory can be associated to a (generalized) YM2 theory in a completely different way, inspired by Matrix String Theory (MST) . This new approach, introduced in Refs. for the case of the cylinder and the torus, and generalized in to arbitrary Riemann surfaces, will be the subject of this talk. It originates quite naturally from quantization of YM2 in a non-propagating gauge, the unitary gauge, but it requires a generalization of YM2 in the sense of . On the other hand, the string interpretation is not related in this case to the expansion in powers of $`\frac{1}{N}`$, and it is valid for any value of $`N`$. The plan of this contribution is the following. In Sec. 2 we review the exact solution of lattice YM2, and its interpretation as a string theory in the large-$`N`$ limit. Sec. 3 contains a general discussion of MST; Sec. 4 discusses the quantization of YM2 in the unitary gauge. In Sec. 5 we introduce the generalized YM2 and its interpretation as a MST. In Sec. 6 we extend the previous considerations to Riemann surfaces with boundaries, while in Sec. 7 we introduce a lattice gauge theory whose partition function coincides with the one of our generalized YM2. Sec. 8 is devoted to some concluding remarks and outlines of future research. ## 2 YM2 ON ARBITRARY RIEMANN SURFACES The exact expression of the partition function of YM2 on a general Riemann surface was first obtained on the lattice by Rusakov , based on previous work by Migdal . It is convenient to consider the lattice theory defined by the heat-kernel action: if the product of the group elements around a plaquette is $`V`$, its Boltzmann weight is $$\mathrm{e}^{S_t(V)}=\underset{R}{}d_R\chi _R(V)\mathrm{e}^{\frac{t}{2}C_R},$$ (1) where the sum is over all the irreducible representations $`R`$ of the gauge group, $`d_R`$ is the dimensionality and $`C_R`$ the quadratic Casimir of the representation, $`\chi _R(V)`$ the character of $`V`$, and $`t=g_{\mathrm{YM}}^2A`$, with $`g_{\mathrm{YM}}`$ the gauge coupling and $`A`$ the area of the plaquette. Notice that $`t`$ is a-dimensional and it measures the area in units of the dimensional coupling constant $`g_{\mathrm{YM}}`$. This action has a unique property: if we integrate over a link variable $`U`$ belonging to two adjoining plaquettes $`A_1`$ and $`A_2`$ we obtain exactly the Boltzmann weight corresponding to the heat kernel action for the resulting plaquette $`A_1A_2`$. In formulae this reads: $$𝑑U\mathrm{e}^{S_{t_1}(V_1U^{})S_{t_2}(UV_2)}=\mathrm{e}^{S_{t_1+t_2}(V_1V_2)}.$$ (2) This property is a simple consequence of the orthogonality of the characters, and it allows to reduce, by successive integrations, any kernel with an arbitrary number of boundaries to a single plaquette with the sides suitably identified, and eventually to compute it exactly by using again well known properties of the characters. On the other hand, the continuum limit can be reached smoothly by adding more and more links to make the lattice spacing smaller and smaller, and the same renormalization invariance (2) insures that the result obtained on the lattice coincides with the one in the continuum. The result of the calculation for a general Riemann surface of genus $`G`$ with an arbitrary number $`p`$ of boundaries is $`K_{G,p}(g_1,\mathrm{},g_p;t)`$ $`={\displaystyle \underset{R}{}}d_R^{22Gp}\chi _R(g_1)\mathrm{}\chi _R(g_p)\mathrm{e}^{\frac{t}{2}C_R},`$ (3) where the $`g_i`$’s are the group elements defined on the boundaries. For $`p=0`$ we obtain the partition function $$Z_G(t)=\underset{R}{}d_R^{22G}\mathrm{e}^{\frac{t}{2}C_R}.$$ (4) From now on we will consider the gauge group $`\mathrm{U}(N)`$ only, whose irreducible representations are labeled by $`N`$ integers $$n_1>n_2>\mathrm{}>n_N.$$ (5) The corresponding dimensions $`d_R`$ and Casimir invariants $`C_R`$ are given by: $`d_R`$ $`=`$ $`{\displaystyle \underset{i>j}{}}(n_in_j),`$ (6) $`C_R`$ $`=`$ $`{\displaystyle \underset{i}{}}n_i^2,`$ (7) so that the partition function reads $`Z_G(N,t)`$ $`=`$ $`{\displaystyle \underset{n_1>n_2>\mathrm{}>n_N}{}}{\displaystyle \underset{i>j}{}}(n_in_j)^{22G}`$ (8) $``$ $`\mathrm{e}^{\frac{t}{2}_in_i^2}.`$ In Refs. the large-$`N`$ limit of Eq. (8) was investigated. The limit was taken, as introduced by ‘t Hooft , by sending $`N\mathrm{}`$ and $`t0`$ while keeping $`\stackrel{~}{t}Nt`$ constant. The result of the analysis is that the partition function (8) in such limit can be interpreted as the theory of a string living in the same two-dimensional target space $`M_G`$ as the original gauge theory. Each string configuration is described by a covering map from a two-dimensional world-sheet $`W_g`$ into $`M_G`$. The admissible string configurations have no folds, so that the area of the world-sheet is an integer multiple of the area of the target manifold. The covering maps are allowed to have branch points. The equivalence of YM2 with the string theory described above is expressed by the following equation<sup>1</sup><sup>1</sup>1We consider here a simplified case where the sum over representations is restricted to “chiral” representations of $`\mathrm{U}(N)`$. Also the so called $`\mathrm{\Omega }`$-point type of singularities, which are present for $`G>1`$, have been neglected. See for details.: $$Z_G(N,t)\begin{array}{c}\\ \\ N\mathrm{}\end{array}\underset{i,g,n}{}N^{22g}\mathrm{e}^{n\stackrel{~}{t}/2}(\stackrel{~}{t})^i\omega _{g,G}^{n,i}.$$ (9) In (9), $`\omega _{g,G}^{n,i}`$ is the number of $`n`$-coverings of the target manifold with genus $`g`$ and $`i`$ branch points. ## 3 MATRIX STRING THEORIES (MST) The interpretation of a generalized YM2 as a string theory that we want to present is inspired to the mechanism through which string states emerge in MST. We begin by briefly introducing a definition of Matrix String Theories sufficiently general to encompass both the original work and our description of YM2. Consider a gauge theory defined on a Riemann surface, and containing at least one field $`F`$ transforming in the adjoint representation of the gauge group: in the case of $`\mathrm{U}(N)`$, $`F`$ is a hermitean $`N\times N`$ matrix. A possible gauge choice consists in diagonalizing $`F`$, that is in conjugating it into an element of the Cartan sub-algebra. This gauge choice is characterized by a Gribov ambiguity related to the Weyl group, namely the group of invariance of the Cartan sub-algebra, which in our case is the symmetric group $`S_N`$: in fact given a hermitian matrix $`F`$ there exist $`N!`$ matrices that diagonalize $`F`$, corresponding to all possible permutations of the eigenvalues. So in order to fix the gauge completely one has to choose in each point one of the $`N!`$ Gribov copies. It is easy to see that in general this cannot be done smoothly on the whole Riemann surface whenever the first homotopy group of the surface is non-trivial. In fact let $`\gamma `$ be a homotopically non-trivial closed path based in a point $`x`$, and choose one of the Gribov copies in $`x`$, say the one, named standard in the following, where $`F=\mathrm{diag}(\lambda _i)`$ with $`\lambda _1(x)\lambda _2(x)\mathrm{}\lambda _N(x)`$. If we diagonalize $`F`$ smoothly along $`\gamma `$ and we return to $`x`$, we will in general end up on a different Gribov copy, related to the standard one by a permutation $`P(\gamma )`$. In this way we can associate a permutation to each homotopy class of closed paths based in $`x`$. It is easy to convince oneself that this defines a group homomorphism $$P:\pi _1(M;x)S_N.$$ (10) We conclude that each configuration of the adjoint field $`F`$ belongs to a topological sector identified by a group homomorphism (10). Consider for example the case of the torus: the homotopy group is generated by two cycles $`a`$ and $`b`$ with the relation $$ab=ba.$$ (11) Therefore to each configuration of the field $`F`$ and to each base point $`x`$ we can associate a topological sector defined by a pair of commuting permutations $`(P_1,P_2)`$; changing the base point $`x`$ results in changing the pair $`(P_1,P_2)`$ into a conjugate pair $`(QP_1Q^1,QP_2Q^1)`$. The homomorphisms (10) are in one-to-one correspondence with the $`N`$-coverings of the surface. This correspondence can be easily visualized by imagining each eigenvalue of $`F`$ as living on one of the $`N`$ sheets of the covering. Therefore we seem to be in a position to conclude that for a general Riemann surface, any theory with a $`\mathrm{U}(N)`$ gauge symmetry and at least one field transforming in the adjoint representation can be written as a sum over coverings of the surface, that is as a string theory. While this is true in principle, it is of real interest only if the resulting string theory is free or weakly coupled, that is if the different sheets of the $`N`$-covering (i.e. the different eigenvalues of $`F`$) do not interact among themselves or interact weakly. In general, however, strong interactions between the eigenvalues of $`F`$ do occur, coming from two different sources: 1. from the functional integration over the other fields of the theory; 2. from the Faddeev-Popov determinant induced by the gauge-fixing. The latter source of interactions is obviously always present, since the Faddeev-Popov determinant is just the Vandermonde determinant of the eigenvalues: $$\mathrm{\Delta }_{\mathrm{FP}}=\mathrm{exp}d^2x\mathrm{\hspace{0.33em}2}\sqrt{g}\underset{i>j}{}\mathrm{log}|\lambda _i\lambda _j|.$$ (12) Therefore, for the gauge theory to be a (weakly coupled) string theory the Vandermonde determinant must be canceled by the integration over other fields. In Ref. , a MST was introduced as a non-perturbative formulation of type-IIA superstring theory. It is described in terms of a supersymmetric, two-dimensional $`\mathrm{U}(N)`$ gauge theory with action $`S`$ $`=`$ $`{\displaystyle }d^2\xi \mathrm{Tr}[D_\mu X^iD^\mu X^i+{\displaystyle \frac{1}{2}}g_\mathrm{s}^2F_{\mu \nu }F^{\mu \nu }`$ (13) $`+`$ $`{\displaystyle \frac{1}{2g_\mathrm{s}^2}}[X^i,X^j]^2i\overline{\theta }\gamma ^\mu D_\mu \theta `$ $``$ $`{\displaystyle \frac{1}{g_\mathrm{s}}}\theta ^T\mathrm{\Gamma }_i[X^i,\theta ]],`$ where $`g_\mathrm{s}`$ is the string coupling, $`X^i`$ $`(i=1,\mathrm{},8)`$ are $`N\times N`$ hermitean matrices representing the transverse space-time coordinates, $`\theta `$ their fermionic superpartners, and $`F`$ is the $`\mathrm{U}(N)`$ field strength. In the limit $`g_\mathrm{s}0`$ all the fields $`X`$ and $`\theta `$ commute and can be diagonalized simultaneously. This diagonalization produces the topological sectors as described above. The cancellation of the Vandermonde determinant is achieved through supersymmetry: the contribution of the fermionic fields $`\theta `$ cancels the one of the $`X`$ fields, so that the strong interaction between the eigenvalues disappears and the theory indeed describes free strings propagating in 8 transverse dimensions. For nonzero $`g_\mathrm{s}`$, weak string interactions of order $`g_\mathrm{s}^2`$ are generated by the non-diagonal terms. ## 4 YM2 IN THE UNITARY GAUGE In this section we will discuss the quantization of YM2 in the unitary gauge, namely in the gauge where the field strength $`F`$ is diagonal. This condition does not fix the gauge completely as it leaves a $`\mathrm{U}(1)^N`$ gauge invariance, so that further gauge fixing conditions for the abelian part have to be introduced separately. The unitary gauge is a non-propagating gauge (the non-diagonal components of the ghost-antighost fields do not propagate) and in four dimensions it makes the theory non-manifestly renormalizable. As we shall see, also in two dimensions this gauge is plagued by divergences that occur when two eigenvalues coincide. However, these can be eliminated by considering a suitable generalized YM2 (in the sense of ). This procedure also eliminates the interaction between different eigenvalues induced by the Vandermonde determinants, thus ensuring that the strings associated to the topological sectors discussed in the previous section are essentially free. The starting point is the first order formalism for YM2 on a genus-$`G`$ Riemann surface: the partition function is $`Z`$ $`=`$ $`{\displaystyle }[dA][dF]\mathrm{exp}[{\displaystyle _{\mathrm{\Sigma }_G}}d\mu V(F)`$ (14) $`+`$ $`\mathrm{i}\mathrm{Tr}{\displaystyle }f(A)F],`$ where $$f(A)=dA+AA,$$ (15) F is an auxiliary field transforming in the adjoint representation of $`\mathrm{U}(N)`$, and $`V(F)`$ is a gauge-invariant potential, depending on the eigenvalues of $`F`$ only. This defines a generalized Yang-Mills theory ; ordinary YM2 is recovered by choosing $$V(F)=\mathrm{Tr}\frac{t}{2}F^2$$ (16) and integrating over $`F`$ to obtain the usual second-order formalism. To determine whether such a theory is a MST, we have to 1. fix the gauge where $`F`$ is diagonal; 2. check whether the Vandermonde determinant coming from the gauge fixing can be canceled by terms coming from other fields. The BRST-invariant action for the unitary gauge is of the form $$S_{\mathrm{BRST}}=S_{\mathrm{Cartan}}+S_{\mathrm{off}\mathrm{diag}},$$ (17) where $`S_{\mathrm{Cartan}}`$ depends only on the eigenvalues of $`F`$ and the diagonal components of the $`A`$ fields: $$S_{\mathrm{Cartan}}=_{\mathrm{\Sigma }_G}[V(\lambda )\mathrm{i}\underset{i=1}{\overset{N}{}}\lambda _idA^{(i)},]$$ (18) while $`S_{\mathrm{off}\mathrm{diag}}`$ is given by $`S_{\mathrm{off}\mathrm{diag}}`$ $`=`$ $`{\displaystyle }d\mu {\displaystyle \underset{i>j}{}}(\lambda _i\lambda _j)[{\displaystyle \frac{1}{2}}ϵ^{ab}\widehat{A}_a^{ij}\widehat{A}_b^{ji}`$ (19) $`+`$ $`\mathrm{i}c^{ij}\overline{c}^{ji}+\mathrm{i}c^{ji}\overline{c}^{ij}],`$ where $$\widehat{A}_a^{ij}E_a^\mu A_\mu ^{ij},$$ (20) with $`E_a^\mu `$ the inverse vierbein; $`c`$ and $`\overline{c}`$ are the ghost and anti-ghost fields. If YM2 is a MST, the integrations over $`\widehat{A}`$ and $`c,\overline{c}`$ must cancel each other. Note that at the classical level $`S_{\mathrm{off}\mathrm{diag}}`$ is invariant under the following transformation, depending on four Grassmann-valued parameters $`\eta `$, $`\zeta `$, $`\xi `$, $`\chi `$: $`\delta \widehat{A}_0`$ $`=`$ $`\mathrm{i}(\eta c+\zeta \overline{c}),`$ (21) $`\delta \widehat{A}_1`$ $`=`$ $`\mathrm{i}(\xi c+\chi \overline{c}),`$ (22) $`\delta c`$ $`=`$ $`\chi \widehat{A}_0+\zeta \widehat{A}_1,`$ (23) $`\delta \overline{c}`$ $`=`$ $`\xi \widehat{A}_0+\eta \widehat{A}_1.`$ (24) This “supersymmetry” does indeed seem to guarantee the cancellation of the contribution of $`S_{\mathrm{off}\mathrm{diag}}`$. However, it is known that this supersymmetry is in general broken by quantum effects, the anomaly being proportional to the scalar curvature of $`\mathrm{\Sigma }_G`$: $`{\displaystyle [dc][d\overline{c}][dA]\mathrm{e}^{S_{\mathrm{off}\mathrm{diag}}}}`$ $`=\mathrm{exp}{\displaystyle \frac{1}{8\pi }}{\displaystyle _{\mathrm{\Sigma }_G}}𝑑\mu R{\displaystyle \underset{i>j}{}}\mathrm{log}|\lambda _i\lambda _j|.`$ (25) This anomaly is essentially due to the fact that for a curved surface (twice) the number of 0-forms (the $`c`$ and $`\overline{c}`$ fields) does not equal the number of 1-forms (the field $`A`$). Inserting this result into the partition function we obtain an effective theory for the diagonal degrees of freedom that exhibit a residual $`\mathrm{U}(1)^N`$ gauge invariance: $`Z`$ $`=`$ $`{\displaystyle }[dA^{(i)}][d\lambda _i]\mathrm{exp}{\displaystyle _{\mathrm{\Sigma }_G}}[\stackrel{~}{V}(\lambda )d\mu `$ (26) $``$ $`\mathrm{i}{\displaystyle \underset{i}{}}\lambda _idA^{(i)}],`$ with $$\stackrel{~}{V}(\lambda )=V(\lambda )\frac{1}{8\pi }_{\mathrm{\Sigma }_G}R\underset{i>j}{}\mathrm{log}|\lambda _i\lambda _j|.$$ (27) The last term at the r.h.s. of (27) is singular when two eigenvalues coincide and the problem of a consistent regularization leading to the standard expression (8) for the partition function is essentially unsolved. According to Ref. , the procedure to recover (8) from (26) involves two steps: 1. The $`\mathrm{U}(1)^N`$ gauge fixing and the functional integration over the $`A^{(i)}`$ fields are performed, and the result is equivalent to replacing all the eigenvalues $`\lambda _i(x)`$ with integer-valued constants $`n_i`$ (the flux of the abelian field strength) and summing over the $`n_i`$’s. One obtains $$Z=\underset{n_1,\mathrm{},n_N}{}\underset{i>j}{}\left(n_in_j\right)^{22G}\mathrm{e}^{V(n)}.$$ (28) 2. The terms with two integers $`n_i`$ coinciding, to be referred to in the following as non-regular terms, are divergent and are essentially discarded by hand. In this way, with $`V(n)=\frac{t}{2}_in_i^2`$ one recovers Eq. (8). Discarding non-regular terms also means discarding the non-trivial topological sectors described in Sec. 3. In fact in a non-trivial topological sector there is at least one homotopically non-trivial closed path that induces a non-trivial permutation on the eigenvalues. As the eigenvalues are constant integers, this implies that at least two eigenvalues coincide. In the following section we propose an alternative approach, that allows to eliminate all divergences while keeping all non-trivial sectors at the price, however, of modifying the coupling of YM2 to gravity. ## 5 GENERALIZED YM2 IS A MATRIX STRING THEORY The divergences of the theory in the unitary gauge can be eliminated, rather than by the ad hoc procedure of Ref. , by a redefinition of the theory, namely by using the arbitrariness of the potential $`V(\lambda )`$ to absorb the divergent term due to the anomaly. In other words since all choices of $`V(\lambda )`$ define a legitimate generalized Yang-Mills theory, it is apparent that there is a specific choice in which the anomaly cancels and one obtains a Matrix String Theory: namely we can choose $`V(\lambda )`$ in such a way that the anomaly term cancels and $`\stackrel{~}{V}(\lambda )`$ is regular when two eigenvalues coincide. For example we can choose $`V(\lambda )`$ such that $$\stackrel{~}{V}(\lambda )=\frac{t}{2}\underset{i}{}\lambda _i^2.$$ (29) With this choice we define a theory whose action coincides with the one of ordinary YM2 on flat surfaces, but differs from it when coupled to two-dimensional gravity (recall that the anomaly (25) is proportional to the curvature of the Riemann surface). The partition function of the generalized YM2 defined by the potential (29) consists of topological sectors which are in one-to-one correspondence with the inequivalent coverings of the target manifold. There is no interaction among different eigenvalues, so the partition function describes a theory of non interacting strings. It is clear from Eq. (26) that a $`\mathrm{U}(1)`$ gauge theory is defined on each connected component of the covering, that is of the world sheet. The product of the partition functions of these $`\mathrm{U}(1)`$ gauge theories gives the Boltzmann weight of the covering. Although a compact expression for the partition function is in general not available we present here the recipe for writing it down for a general Riemann surface without boundaries (the case of a surface with boundaries will be discussed in Sec. 6). It is convenient to work directly on the generating function of the partition functions for arbitrary $`N`$, namely the grand-canonical partition function. This is defined as $$Z(G,q)=\underset{N=0}{\overset{\mathrm{}}{}}Z_N(G)q^N,$$ (30) where $`Z_N(G)`$ is the $`\mathrm{U}(N)`$ partition function. For a given Riemann surface $`\mathrm{\Sigma }_G`$ of genus $`G`$ one proceeds as follows: 1. Write the grand-canonical partition function of unbranched coverings of $`\mathrm{\Sigma }_G`$: $$Z^{(\mathrm{cov})}(G,q)=\underset{k=0}{\overset{\mathrm{}}{}}Z_k^{(\mathrm{cov})}(G)q^k,$$ (31) where $`Z_k^{(\mathrm{cov})}(G)`$ is the number of inequivalent $`k`$-coverings of $`\mathrm{\Sigma }_G`$. $`Z^{(\mathrm{cov})}(G,q)`$ has been studied in the literature and is known in closed form. 2. Take the logarithm of $`Z^{(\mathrm{cov})}(G,q)`$ to obtain the free energy $`F^{(\mathrm{cov})}(G,q)`$ $`=`$ $`\mathrm{log}Z^{(\mathrm{cov})}(G,q)`$ (32) $`=`$ $`{\displaystyle \underset{k=0}{\overset{\mathrm{}}{}}}F_k^{(\mathrm{cov})}(G)q^k,`$ where $`F_k^{(\mathrm{cov})}(G)`$ is the number of connected $`k`$-coverings of $`\mathrm{\Sigma }_G`$. No closed expression is known for $`F^{(\mathrm{cov})}(G,q)`$. 3. Associate to each connected covering the partition function of the $`\mathrm{U}(1)`$ gauge theory living on it, to obtain the free energy of the generalized YM2: $$F(G,q,t)=\underset{k=0}{\overset{\mathrm{}}{}}F_k^{(\mathrm{cov})}(G)Z_{\mathrm{U}(1)}(k,t)q^k.$$ (33) Here $`Z_{\mathrm{U}(1)}(k,t)`$ is the partition function of a (generalized) $`\mathrm{U}(1)`$ gauge theory on a Riemann surface of area $`kt`$ (in units of the gauge coupling constant). $`Z_{\mathrm{U}(1)}(k,t)`$ does not depend on the genus and for the quadratic potential (29) it reads<sup>2</sup><sup>2</sup>2The generalization to arbitrary potentials is straightforward: $$Z_{\mathrm{U}(1)}(k,t)=\underset{n=\mathrm{}}{\overset{\mathrm{}}{}}\mathrm{e}^{\frac{1}{2}ktn^2}.$$ (34) Therefore $$F(G,q,t)=\underset{n=\mathrm{}}{\overset{\mathrm{}}{}}F^{(\mathrm{cov})}(G,\mathrm{e}^{\frac{1}{2}tn^2}q).$$ (35) 4. The gauge theory partition function is then $`Z(G,q,t)`$ $`=`$ $`\mathrm{e}^{F(G,q,t)}={\displaystyle \underset{N=0}{\overset{\mathrm{}}{}}}Z_N(t,G)q^N`$ (36) $`=`$ $`{\displaystyle \underset{n}{}}Z^{(\mathrm{cov})}(G,\mathrm{e}^{\frac{1}{2}tn^2}q).`$ As an example, we will treat explicitly the case of the torus. This is the only case (except for the trivial one of the sphere) where a closed expression is available for the number of connected coverings. As discussed in Sec. 3, the coverings of the torus are in one-to-one correspondence with the pairs of commuting permutations. The number of such pairs is $`N!p(N)`$ where $`p(N)`$ is the number of partitions of $`N`$. Therefore there are $`p(N)`$ inequivalent coverings (that is, not counting as different two coverings that can be obtained from each other by simple relabeling of the sheets) and we have: $`Z^{(\mathrm{cov})}(G=1,q)`$ $`=`$ $`{\displaystyle \underset{N=0}{\overset{\mathrm{}}{}}}p(N)q^N`$ (37) $`=`$ $`{\displaystyle \underset{k=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{1}{1q^k}}.`$ For the connected coverings we have $$F^{(\mathrm{cov})}(G=1,q)=\underset{N=1}{\overset{\mathrm{}}{}}\underset{r|N}{}\frac{1}{r}q^N,$$ (38) where the second sum is extended to all divisors of $`N`$ (including $`N`$). Therefore we obtain for the YM2 free energy the expression $$F(G=1,q,t)=\underset{n=\mathrm{}}{\overset{\mathrm{}}{}}\underset{N=1}{\overset{\mathrm{}}{}}\underset{r|N}{}\frac{1}{r}\mathrm{e}^{\frac{1}{2}Ntn^2}q^N$$ (39) and for the partition function $$Z(G=1,q,t)=\underset{n=\mathrm{}}{\overset{\mathrm{}}{}}\underset{k=1}{\overset{\mathrm{}}{}}\frac{1}{1q^k\mathrm{e}^{\frac{1}{2}ktn^2}}.$$ (40) This result does not coincide with the expression (28) of the partition function for $`G=1`$. To recover the “usual” partition function on the torus, one has to consider only the topological sectors corresponding to pairs of commuting permutations of the special form $`(1,P)`$, weighted with a factor $`(1)^{n_c}`$ where $`n_c`$ is the number of connected components of the covering. One can check that this prescription effectively cancels all non-regular terms from Eq. (28). Therefore for the case of the torus, where the Riemann curvature vanishes and our generalized YM2 coincides with the usual one at the classical level, the inclusion of all the topological sectors leads to a different quantum theory which is richer in structure and closely related to the MST of Ref. , as shown in Refs. . ## 6 SURFACES WITH BOUNDARIES Consider now a Riemann surface $`\mathrm{\Sigma }_{G,p}`$ of genus $`G`$ with $`p`$ boundaries. The arguments of the preceding sections hold essentially unaltered, and the generalized YM2 can be written as a theory of $`N`$-coverings of $`\mathrm{\Sigma }_{G,p}`$ with a $`\mathrm{U}(1)`$ gauge theory living on each connected component of the covering. The natural quantity to be studied is the kernel that will depend on $`p`$ states defined on each boundary. For any given boundary, a state is completely specified by assigning: 1. A conjugacy class of permutations, defined as the conjugacy class of the permutation $`P(\gamma )`$ where $`\gamma `$ is any closed curve that can be continuously deformed into the boundary. It is given in terms of $`N`$ integers $`r_l`$ $`(l=1,\mathrm{},N)`$ satisfying $`_llr_l=N`$ where $`r_l`$ is the number of cycles of length $`l`$ in $`P(\gamma )`$. 2. A $`\mathrm{U}(1)`$ holonomy associated to each cycle of the conjugacy class defined above. The procedure to construct the kernels follows closely the one described above for surfaces without boundaries: one constructs the free energy of the coverings of $`\mathrm{\Sigma }_{G,p}`$ and associates to each connected $`k`$-covering a $`\mathrm{U}(1)`$ kernel $$K_{\mathrm{U}(1)}(kt,\varphi )=\underset{n=\mathrm{}}{\overset{\mathrm{}}{}}\mathrm{e}^{\frac{1}{2}ktn^2+\mathrm{i}n\varphi }.$$ (41) Exponentiation then produces the generalized YM2 kernel. Details can be found in Ref. . One must check the consistency of the theory, namely that the set of kernels constructed in this way produces consistent results when the surfaces are cut and sewn to produce new ones. This property is obviously true for the pure theory of coverings and it can be proved that the introduction of the $`\mathrm{U}(1)`$ gauge theory on each connected world sheet does not spoil it. ## 7 LATTICE EFFECTIVE THEORY We have seen in the previous sections that, after integration over the non-diagonal components of the fields, our generalized YM2 theory is described by an effective theory of coverings with a $`\mathrm{U}(1)`$ gauge theory on each connected world sheet. In this section we show that this effective theory is equivalent to a lattice theory whose gauge group is the semi-direct product of $`S_N`$ and $`\mathrm{U}(1)^N`$. Consider first a surface with the topology of a disc. Since its fundamental group is trivial, the only $`N`$-covering consists of $`N`$ disjoint copies of the disc: all closed paths are mapped into the identity by the homomorphism (10). From the discussion of the previous section it follows that the state on the boundary is given in terms of $`N`$ $`\mathrm{U}(1)`$ invariant angles $`\varphi _i`$. The kernel is then $$K((P,\varphi ),t)=\delta (P)\underset{n_i}{}\mathrm{e}^{_{i=1}^N\left(\mathrm{i}n_i\varphi _itv(n_i)\right)}.$$ (42) Consider now the subgroup $`𝒢_N`$ of $`\mathrm{U}(N)`$ given by the matrices of the form $$(P,\varphi )=\mathrm{diag}(\mathrm{e}^{\mathrm{i}\varphi _1},\mathrm{},\mathrm{e}^{\mathrm{i}\varphi _N})P,$$ (43) where the $`\varphi _i`$’s are real and $`P`$ is a permutation matrix: $$P_{ij}=\delta _{iP(j)}.$$ (44) The group product reads $$(P,\varphi )(Q,\theta )=(PQ,\varphi +P\theta ),$$ (45) which shows that $`𝒢_N`$ is a semi-direct product of $`S_N`$ and $`\mathrm{U}(1)^N`$. The kernel (42) is invariant under gauge transformations belonging to $`𝒢_N`$: $$K((Q,\theta )(P,\varphi )(Q,\theta )^1,t)=K((P,\varphi ),t).$$ (46) Moreover it shares with the heat-kernel action (1) of lattice gauge theory the invariance under renormalization group transformations: $`{\displaystyle \underset{QS_N}{}}{\displaystyle \underset{i}{}\frac{d\theta _i}{2\pi }K((P_1,\varphi _1)(Q,\theta ),t_1)}`$ $`K((Q,\theta )^1(P_2,\varphi _2),t_2)`$ $`=K((P_1,\varphi _1)(P_2,\varphi _2),t_1+t_2).`$ (47) The kernel (42) can thus be used as a heath-kernel type of action to construct a lattice gauge theory for the group $`𝒢_N`$, with partition function $$Z_{\mathrm{latt}}=\underset{\alpha }{}\underset{P_\alpha }{}_0^{2\pi }𝑑\varphi _\alpha \underset{\mathrm{pl}}{}K((P_{\mathrm{pl}},\varphi _{\mathrm{pl}}),t),$$ (48) where $`\alpha `$ enumerates the links in the lattice, and the product of all the elements $`(P_\alpha ,\varphi _\alpha )`$ of $`𝒢_N`$ along a plaquette defines the plaquette variable $`(P_{\mathrm{pl}},\varphi _{\mathrm{pl}})`$. In complete analogy with the solution of YM2 obtained in , one can solve exactly the theory by reducing the surface to a single plaquette with suitably identified links. As shown in , the resulting expression coincides with the one that is obtained, via the steps discussed in the previous Section, by including all the topological sectors in the MST corresponding to our generalized YM2 theory. ## 8 CONCLUSIONS AND FURTHER DEVELOPMENTS Quantization of YM2 in a non-propagating gauge, such as the unitary gauge, leads to non-manifestly renormalizable divergences as well as to the appearance of a structure of topological sectors, associated to the $`N`$ coverings of the target manifold ; this strongly suggests the presence of an underlying string theory. In order to recover the standard partition function obtained with other gauge choices, the divergences have to be removed more or less by hand, thus suppressing the topological sectors even in the case of torus where they are not related to any divergence. However we found that the divergences can be removed, and the topological sectors maintained, by considering a generalized theory whose action coincides with standard one for vanishing Riemann curvature. Indeed one can say that the generalized theory differs from the conventional one for a non-standard coupling with gravity. In this approach the string theory does not emerge from the large $`N`$ limit, but directly for any value of $`N`$ from the topological structure of the theory as in MST. It is not a chance that for the case of the torus our partition function is closely related to the one of the full MST. These considerations suggest some lines along which future research might develop. Two of them seem the most natural ones: 1. The string theory considered so far is a theory of free strings. It is natural to introduce interactions by including branched coverings, that is the possibility for strings to split and join. A branch point can be seen as a boundary where all the $`\mathrm{U}(1)`$ invariant angles are set to zero. In this sense our treatment of surfaces with boundaries includes branch points as well. However, the most interesting case is the one in which a continuum distribution of branch points is introduced, as was done in Refs. for the pure theory of coverings. 2. Other non-propagating gauges would require a different tuning of the potential to eliminate the divergences and may lead to more general theories. For instance a gauge choice where, for $`N=LM`$, the field strength is block diagonal, with blocks of size $`M\times M`$, would be equivalent to a theory of $`L`$-coverings with a non abelian $`U(M)`$ gauge theory on the world sheet. This formulation would include both standard YM2 ($`L=1`$) and our generalized theory ($`L=N`$) as limiting cases. As a string theory emerges in both cases through completely different mechanisms, some investigation will be needed to ascertain whether a deeper relation exists between these two different formulations.
warning/0001/cond-mat0001200.html
ar5iv
text
# Time Dependent Current Oscillations Through a Quantum Dot ## I Abstract Time dependent phenomena associated to charge transport along a quantum dot in the charge quantization regime is studied. Superimposed to the Coulomb blockade behaviour the current has novel non-linear properties. Together with static multistabilities in the negative resistance region of the I-V characteristic curve, strong correlations at the dot give rise to self-sustained current and charge oscillations. Their properties depend upon the parameters of the quantum dot and the external applied voltages. PACS numbers: 73.40.Gk, 85.30.Mn ## II Introduction Electron correlation effects on the transport properties of a quantum dot (QD) under the influence of an external potential have been the subject of many experimental and theoretical studies in recent years. This device although possesses thousands of atoms behaves as if it were an artificial atom due to its discrete energy spectrum. The enormous advantages of this artificial atom are that its properties can be modified continuously by adjusting the external potentials applied to it . Unlike a real atom it can be experimentally isolated and its transport properties obtained studying the flow of charges through it. The quantization of charge and energy plays the major role in the transport properties of a QD. The conductance exhibits oscillations as a function of the external potential which can be explained in terms of a transport mechanism governed by single-electron tunneling and Coulomb blockade effect due to the e-e interaction inside the dot. It is well established that the Coulomb interaction produces non-linear effects in 3D double barrier heterostructure, observed as intrinsic static bistability in the negative differential resistance region of the I-V characteristic curve of the device. This property has been explained as an electrostatic effect due to the e-e interaction among the charges inside the well. Besides this static behavior, theoretical studies have proposed that the Coulomb interaction can produce time dependent oscillations. More recently current oscillations promoted by an external magnetic field applied in the direction of the current were predicted. These last results showed that the system bifurcates as the field is increased and may transit to chaos at large enough fields. Recently very interesting experimental evidences of the existence of these oscillations induced by an external magnetic field, have been given. Although Coulomb interaction phenomena have been a matter of great concern in the last years, to the best of our knowledge, there have not been experimental studies reporting neither static bistabilities nor current and charge oscillations through a QD. In the search for stationary behavior, we were able to theoretically predict a bistability similar to the one that appears in the I-V characteristic curve of a 3D double barrier heterostructure. We obtained as well evidences of the non-existance of stationary solutions in some regions of the parameter space for the currents flowing through the QD. In this communication letter we address the study of currents going through a QD connected to leads under the effect of an external potential. We report the existence of static hysteresis behavior of the current in the negative resistance region. What is more important, we find novel time dependent phenomena which appear as self-sustained current and charge oscillations with properties depending on the internal parameters of the system and the external applied potentials. These time dependent phenomena are derived from the highly correlated electrons inside the QD. It has been shown that the physics associated to a QD connected to two leads can be readily understood in terms of a single magnetic impurity Anderson Hamiltonian, where the impurity is the quantum dot. The leads can be represented by a 1D tight-binding Hamiltonian connecting the dot to particle reservoirs characterized by Fermi levels $`E_l`$ and $`E_r`$. The difference $`E_lE_r`$ gives the bias voltage applied to the system when connected to a battery. The Hamiltonian is given by $`H`$ $`=`$ $`{\displaystyle \underset{i\sigma }{}}ϵ_in_{i\sigma }+v{\displaystyle \underset{<ij>\sigma }{}}c_{i\sigma }^{}c_{j\sigma }+(ϵ_o+V_g){\displaystyle \underset{\sigma }{}}n_{o\sigma }+Un_on_o.`$ (1) The diagonal matrix element $`ϵ_i`$ is site dependent because it takes into account the potential profile and the two barriers. The gate voltage $`V_g`$ is supposed to be effective only on the QD. Since we are not interested in very low temperature physics, as for instance the Kondo effect, we use a decoupling procedure to solve the equation of motion for the one particle operator. Although this is a simple approximation, it is adequate to treat high correlations at the dot, responsible for Coulomb blockade phenomena. The non-linear behavior stems from the dependence of the solution upon the self-consistent calculated charge content of the dot. As we are not interested in studying magnetic properties, we omit the subindex $`\sigma `$ corresponding to the spin. Using the set of Wannier functions localized at site i, $`\varphi _i(r)`$, we can write the time dependent state $`\phi _k(r,t)`$ as, $$\phi _k(r,t)=\underset{i}{}a_i^k(t)\varphi _i(r)$$ (2) The probability amplitude for an electron to be at site $`j`$ in a time dependent state $`\phi _k(r,t)`$ is given by the equation of motion, $$i\mathrm{}\frac{da_j^k}{dt}=ϵ_ja_j^k+v(a_{j1}^k+a_{j+1}^k)(i0)$$ (3) $$i\mathrm{}\frac{da_o^{k,\alpha }}{dt}=(ϵ_o^\alpha +V_g)a_o^{k,\alpha }+n^\alpha v(a_1^k+a_1^k)$$ (4) where $`\alpha =\pm `$; $`n^+=n`$; $`n^{}=1n`$; $`ϵ_o^+=U`$; $`ϵ_o^{}=0`$ and $`n`$ is the number of electrons per spin at the dot that is calculated from the probability amplitudes by the equation, $$n=2_0^{k_f}a_o^k^2𝑑k$$ (5) where $`a_o^k=a_o^{k,+}+a_o^{k,}`$ and $`k_f`$ is the Fermi wave vector of the emitter side. We have assumed for simplicity that the collector is not doped so that the charge that goes towards the emitter is restricted to the electrons that have been reflected on the barriers. The time dependent equations are solved by discretizing equations (3) and (4), and using a half-implicit numerical method, which is second-order accurate and unitary. The solution of these equations require the knowledge of the wave function at $`\mathrm{}`$ and $`+\mathrm{}`$. We suppose that an incident electron of wave vector k is described by the function, $$a_j^k=(Ie^{ikx_j}+R_je^{ikx_j})e^{iϵt/\mathrm{}}(x_jL)$$ (6) $$a_j^k=T_je^{ik^{}x_j}e^{iϵt/\mathrm{}}(x_jL)$$ (7) where the system is explicitly defined between $`L`$ and $`L`$ and $`k^{}`$ is the wave vector of the transmitted particle $`k^{}=\sqrt{(}\frac{2m(E_lE_r)}{\mathrm{}^2}+k^2)`$. The incident amplitude I is assumed to be spatially constant. Instead, far from the barriers, the envelope function of the reflected and transmitted waves $`R_j`$ and $`T_j`$ are supposed to be weakly dependent on site $`j`$. This permits to restrict the dependence to the linear term, which results to be an adequate approximation provided the time step taken to discretize equations (3) and (4) is less than certain limit value that depends upon the parametrization of the system. A maximum value of $`0,3fs`$ was adequate to guarantee numerical stability up to 200 ps. To eliminate spurious reflections at the boundary it was necessary to take systems of the order of $`2L=400`$ sites. In the numerical procedure, the Wannier intensities obtained for one bias are used as the starting point for the next bias step. Once the Wannier intensities are known the current is calculated from, $$J_j2_o^{k_f}[Im(a_j^k^{}(a_{j+1}^ka_j^k)]dk$$ (8) It is important to emphasize that the system we study has to be in the Coulomb blockade regime. Charge quantization is obtained satisfying the relation $`\frac{t^2}{W}<U`$ where $`t^{}`$ is an effective coupling constant between the QD and the leads controlled by the barriers parameters and W is the leads bandwidth. In this case the energy spacing U between the states with $`N`$ and $`N+1`$ electrons is greater than the width of the resonant levels within the dot. We study the system defined by an emitter barrier of $`1.4eV`$ and a collector barrier of $`2.4eV`$ both of a thickness of 5 lattice parameters and an electronic repulsion U = 20meV. This configuration represents a GaAs structure in the charge quantization regime. A similar effective coupling constant can be obtained assuming lower and thicker barriers. We have chosen the case with less number of sites in order to reduce the computational effort. In order to enhance the charge content of the dot and consequently the non-linear effects the second barrier is taken to be higher than the first. The Fermi level lies 30 meV above the bottom of the emitter conduction band. With these parameters and zero gate voltage and bias the localized level at the dot lies well above the Fermi level, so that the system is completely out of resonance. For small values of the gate potential equations (3) and (4) have a stationary solution after a transient, for the whole interval of bias voltage $`V`$, which drives the system into and off resonance. Increasing the gate potential the system enters into a completely different regime. As the bias voltage required to drive the system into resonance is small, because the resonant level is now near the Fermi level, the second barrier is not significantly reduced by the bias voltage and is able to trap a large amount of electronic charge inside the dot. As a consequence, the non-linear effect is enhanced giving rise to a region in the I-V characteristic curve where there is no stationary solution. In this parameter region the current has self-sustained oscillations while for greater values of the bias voltage the system possesses complex electrostatic multistabilities which can be obtained, depending upon the initial condition, augmenting or diminishing the bias voltage. These static multistabilities are very well known phenomena of 3D systems, although they generally assume a bistable structure. The situation for the QD described is illustrated in Fig. 1 where the bubble shows the maximum and minimum of the current oscillations, while for a greater value of the bias voltage the electrostatic instabilities are clearly represented. The non-stationary solutions lie outside the static multistable region for small values of $`V_g`$. However, for larger values of $`V_g`$, there is a superposition of the regions where both effects are active. The static large multistability is related to the drop of the resonant level below the bottom of the conduction band, while the dynamic oscillations appear when the resonant level aligns with the Fermi level. The size and location of the oscillating behavior in the I-V space depends upon the gate voltage. Its size increases with $`V_g`$. The non-linearities are enhanced by the gate potential, constituting an external tunable parameter through which the oscillating phenomena can be controlled. We have taken different values of the bias $`V_g`$ to study explicitly the time dependent current going along the device. For $`V_g=0.2`$ volts the system possesses a stationary solution for all values of the bias voltage. For a greater value of the gate potential $`V_g=0.86`$ volts and $`V=5.6`$ volts, we show the current and the charge as a function of time in three different locations of the system: in the emitter and collector side before and after the barriers, in Fig. 2b and, inside the dot, in Fig. 2c. We see that after a transient of some ps, the system begins to oscillate in a self-sustained way. It is important to note the different behaviors of the oscillations depending upon the site where the current is evaluated. The oscillations are large in the region before the emitter barrier and inside the well, while after the charge has passed through the device the more opaque collector barrier damps the oscillation in a very significant way. Although not perfectly periodic, the oscillations have some regularity. The frequency spectrum is continuous with the predominance of two and some times three well define frequency regions. The origin of this behavior is not completely clear to us. However, the existence of more than one frequency region has to be attributed to the amount of non-linearity in the system. A less non-linear situation, obtained for instance in the near vicinity of the bifurcation point where time dependent behavior appears in the I-V characteristic curve of Fig. 1, shows small amplitude oscillating currents of almost only one frequency. Very remarkably, superimposed to the oscillating currents, we exhibit in the inset of Fig. 2c very high frequency oscillations greatly enhanced inside the well and almost non-existent outside it. We attribute them to the electrons going back and forth inside the well with a period corresponding to the time taken by the electrons at the neighborhood of the Fermi energy being successively reflected by the barriers. Changing the gate potential the dot enters into resonance when its local level aligns with the Fermi level. The current increases abruptly and, due to the trapping effect of the collector barrier, the dot charge per spin $`n`$ assumes its maximum value after a time. As a consequence, the strength of the resonant level, which due to many body effects depends upon the dot charge as $`(1n)`$, diminishes and the flow of electrons going into the dot reduces. As time evolves, the total amount of charge inside the dot drops because the leaking of charge through the second barrier becomes greater than its entrance through the first one. The reduction of charge increases the strength of the resonant level, which increases the flowing current starting a new cycle. As we have numerically verified, for a configuration outside the Coulomb blockade regime, by diminishing for instance the first barrier, the oscillations are damped out and the system reaches a stationary regime. In this case, as the system is driven into resonance, the current increases and the dot charge reacts almost instantaneously with no significant lag time between them. To generate self-sustained current oscillations the system has to be in a parameter region where there is no dynamic accommodation between the current and the charge inside the dot. It requires an abrupt entrance in resonance when the external potential is modified. This is the case of the charge quantization regime when the channel for the electron to go through is provided by the very sharp resonant peak inside the dot. It is interesting to compare the current and the charge oscillations inside the well in this case. Although the frequencies and the general properties of both oscillating quantities are the same, there is a clear lag time $`\tau `$ of the dot charge in relation to the current. The dephasing between charge and current which disappears with more transparent barriers is illustrated in Fig. 2a. It is clear that the lag $`\tau `$ depends upon the time scale controlling the entrance of the charge into the resonant state, which is called the buildup time. For opaque emitter barriers, as in our case, the buildup time is found to be essentially independent of the barrier parameters. The buildup process is essentially determined by the spatial distribution of the extra charge in conditions to enter into the well. This is slightly different in each oscillation, creating a rather irregular lag among the two quantities as time goes along, as exhibited in Fig. 2a. Due to computer time limitations it was not possible to develope a systematic investigation of the phenomenon. However, we have studied the system modifying some of its parameters in order to give support to the interpretation above. There is a clear tendency of the oscillation mean value frequency to increase reducing the opacity of the barriers although it is more clearly dependent upon the second barrier. This is consistent with the interpretation given above as it is the collector barrier that controls the leaking time of the charge when the its entrance is abruptly reduced by non-linearity. In order to study the stability of the self-sustained oscillations we have calculated, for one particular case, the time evolution of the system up to a value of $`200`$ps. The oscillations maintain their amplitude and regular shape. We have taken a charged and a discharged dot as two different initial conditions. The first condition creates oscillations immediately after the circuit is switched on while in the second the system goes through a transient during which the dot is charged, as shown in Fig. 3. Although the oscillations have a similar behavior shifting one relative to the other by this charging time, there are differences which in fact increase in time as shown in the inset of Fig. 3. In certain regions of the parameter space the oscillations result to be very sensitive to initial conditions, suggesting the existence of chaotic phenomena. We have done a time dependent study of the current circulating through a QD in the Coulomb blockade regime. We conclude that together with complex static mutistable behavior, non linearity produced by highly correlated electrons in the dot give rise to charge and current time dependent oscillations. They can be controlled by changing the parameters defining the system and the various external potentials applied. A complete characterization of the time dependent oscillations would require a great computational effort, which could be justified only in the case of an experimental verification of the phenomena theoretically predicted in this letter. Work supported by FONDECYT grants 1980225 and 1990443, CONICYT,Catedra Presidencial en Ciencias (Chile),Fundación Antorchas/Vitae/Andes grant A-13562/1-3 and the brazilians agencies FINEP and CNPq.
warning/0001/astro-ph0001108.html
ar5iv
text
# Micro & strong lensing with the Square Kilometer Array: The mass–function of compact objects in high–redshift galaxies ## 1 Introduction Gravitational lens (GL) systems offer a versatile tool to study a range of cosmological and astrophysical problems. One of most puzzling and difficult problems to solve in all of astrophysics, is that of the distribution and nature of dark–matter. Paczynski suggested to search for one of the dark–matter candidates – i.e. massive compact objects, possibly stellar remnants or primordial black holes – using their gravitational lensing effect on background stars. Many so called microlensing surveys have been undertaken since then, with varying success . One major disadvantage of these Galactic microlensing surveys is the low microlensing optical depth ($`\tau `$$``$$`10^6`$) . In multiply–imaged GL systems $`\tau `$$``$1, making them much more suitable to search for microlensing variability (e.g. ). The major disadvantage in this case is the long variability time scale, which can amount to many years. This is especially true if the lens galaxy is at a high redshift and the velocity of the microlensing magnification pattern is dominated by the velocity of the compact objects with respect to the line–of–sight to the stationary optical source. What about microlensing in the radio? Most sources are much more extended in the radio than in the optical. It is therefore often believed that they are nearly unaffected by microlensing. However, many flat–spectrum radio sources contain jet–components moving with near or superluminal velocity. These components can be as small as several $`\mu `$as at the mJy level, to sub–$`\mu `$as at the $`\mu `$Jy level. This is comparable or even smaller than the angular scale over which the microlensing magnification pattern changes significantly. Thus, at these levels radio microlensing can become important. In Sections 2–4, we will illustrate the above with the CLASS GL B1600+434. In Sect.5, we will show how the Square Kilometer Array can improve this in the future. ## 2 The edge-on spiral lens system CLASS B1600+434 The gravitational lens B1600+434 consists of a 1.4–arcsec double, lensed by a foreground edge–on spiral galaxy (Fig.1; ), The lens galaxy and source are at redshifts of 0.41 and 1.59, respectively . B1600+434 was discovered in the Cosmic Lens All–Sky Survey (CLASS), whose mission is to find multiply imaged flat–spectrum radio sources. Early VLA and WSRT radio observations already indicated that the source was variable and therefore suitable for determining a time delay . In the coming paragraphs, we shortly describe constraints on the mass model of the lens galaxy and the time delay between the lens images. Subsequently, we show that most of the short–term variability found in the radio light curves of the lens images is of external origin. Mass models: In a range of mass models to describe the edge–on disk galaxy (Fig.1) were presented, with the aim of constraining the dark–matter halo around the lens galaxy. A lower limit of $`q_\mathrm{h}`$0.5 was found for the oblateness of the dark–matter halo. Moreover, at least half of the mass inside the Einstein radius must be contributed by the dark–matter halo, i.e. the maximum–disk hypothesis does not apply for this galaxy. Assuming an isothermal halo model, a time delay of about 54$`h_{50}^1`$ days was predicted in a flat FRW universe with $`\mathrm{\Omega }_\mathrm{m}=1`$. The steeper MHP mass model for the dark matter halo predict a time delay of about 70$`h_{50}^1`$ days. Once the time delay between the lens images has been determined, constraints on the Hubble parameter can be set. However, more interesting, once the Hubble parameter can be determined from several other JVAS and CLASS lens delays, as well as other methods, the problem can be reversed and constraints can be put on the mass profile of the dark–matter halo around the edge–on spiral lens galaxy . Time-Delay: A monitoring campaign was initiated in Febr. 1998 with the VLA in A– and B–arrays at 8.5–GHZ . B1600+434 was observed on average every 3.3 days. The radio maps have S/N–ratios of $``$300 and a resolution of typically 0.2 (0.7) arcsec in A (B) array (Fig.1). The final error on the calibrated flux densities of the lens images is 0.7–0.8%, dominated by the uncertainties in the flux–density calibration. The normalized calibrated light curves of both images are shown in Fig.2. A 15–20% long–term decrease in flux density of both images is seen over the observing period of eight months, most likely intrinsic source variability . One also sees short term (days to weeks) variability of up to 10% peak–to–peak in image A, whereas the short term variability in image B is significantly less (Fig.2). We determined the time delay by scaling the observed light–curve B by the instrinsic flux–density ratio (1.212) and shifting it back in time, until the dispersion between the two image light curves minimized. We find a time delay of $`\mathrm{\Delta }t_{\mathrm{B}\mathrm{A}}`$=$`47_9^{+12}`$ days (95% confidence; ). The statistical error was determined from Monte–Carlo simulations and a maximum systematic error of $``$8/+7 days was estimated. Note that this method would have yielded a delay even if the source had not shown any short–term variability! External variability: Once the time delay and flux–density ratio have been determined, the light curves can be subtracted to see if any significant variability is left. In Fig.2 the normalized difference light curve for B1600+434 is shown, using a time delay of 47 days and a flux-density ratio of 1.212. The difference light curve has an rms scatter of 2.8%, which is inconsistent with a flat difference light curve at the 14.6–$`\sigma `$ confidence level . The individual normalized light curves (i.e. the light curves divided by a linear fit) of images A and B (Fig.2) have rms variabilities of 2.8% and 1.6%, respectively. Both are significantly above the 0.7–0.8%, expected on the basis of measurement errors. ## 3 Radio Microlensing by MACHOs in B1600+434? Below we investigate both scintillation and microlensing, as plausible causes of the external variability in the lens images of B1600+434 . Scintillation: Scintillation, caused by the Galactic ionized ISM, can explain both large amplitude variability in very compact extra–galactic radio sources and lower amplitude ‘flicker’ of more extended sources (e.g. ). In B1600+434–A and B, we observe several percent rms variability, possibly consistent with ‘flickering’. The time scale is harder to quantify, but it appears that we see variability time scales of several days up to several weeks (Fig.2). Let us summarize the arguments that argue against scintillation: (i) The modulation index between the lens images differs, even though one looks at the same background source. This requires significantly different properties of the Galactic ionized ISM over scales of 1.4<sup>′′</sup>. (ii) Ongoing multi–frequency VLA and WSRT show a decrease in the modulation index with wavelength (Fig.3), whereas the opposite is expected for scintillation. (iii) A modulation index of a few percent corresponds to a variability time scale of less than 1–2 days for weak and strong refractive scattering, whereas variability over much longer scales seems to be present. Although the difference in modulation index can be explained by considerable scatter–broadening of image B in the lens galaxy, all evidence put together build a considerable case against scintillation. Microlensing: There are several good reasons why microlensing is a good explanation for the external variability that is observed in B1600+434: (i) Because B1600+434 is multiply imaged, the lens images pass through a foreground lens galaxy with microlensing optical depths near unity. Combined with the fact that many core–dominated flat spectrum radio source have superluminal jet–structure (e.g. ), makes it probable that microlensing variability is occurring at some level on time scales of weeks to months. (ii) From microlensing simulations , we find that a core plus a single superluminal knot can explain the rms and time scale of variability in the lens images. For image A, however, we require an average mass of compact objects $``$0.5–M to explain its significantly higher rms variability. More complex jets, with multiple components, would of course show reduced modulations, but this can then be compensated by making them somewhat smaller (i.e. a larger Doppler boosting). We have tested this, by replacing the simple source with a real jet–structure (i.e. 3C120), and shown that also more complex sources give observable microlensing variations in their light curves. (iii) The apparent decrease in rms variability (i.e. modulation index) with wavelength appears in agreement with a microlensed synchrotron self–absorbed source that grows proportional with wavelength (Fig.3). The current observations are all in agreement with the microlensing hypothesis, even though some of the short–term variability maybe due to scintillation . In fact, objects compact enough to be microlensed should show scintillation at some level. In the case of microlensing, caustic crossings can brighten objects orders of magnitude more than scintillation if the emitting region is significantly smaller than the Fresnel scale of several $`\mu `$as at 8.5–GHz. In this case we expect variability to be dominated by microlensing. ## 4 B1600+434: Thus far… We have presented the first unambiguous case of external variability of an extra–galactic radio source, the CLASS gravitational lens B1600+434 . Several lines of evidence indicate that it is not dominated by scintillation, but by microlensing (in the lens galaxy) of a superluminal source. This requires a population of MACHOs in the halo around the edge-on disk galaxy with masses $``$0.5–M. The discovery of a population of these objects might well indicate that its dark–matter halo mostly consists of stellar remnants. Maybe these objects are the (white) dwarfs as seen in the Hubble Deep Field . Our ongoing multi–frequency WSRT (Fig.3) and VLA observing campaigns should further tighten the constraints on this tantalizing possibility. ## 5 Future possibilities with the Square Kilometer Array Above we have shown an example of the current possibilities of studying strong and microlensing in the radio. Preferably, one would like to monitor these systems simultaneously at many different frequencies and with much higher sensitivity. Although, we have started a multi-frequency campaign of B1600+434 with the VLA, we are working at the limits of present day technology. The Square Kilometer Array (SKA) has sensitivities of sub-$`\mu `$Jy levels after a few minutes of integration, a resolution of $``$0.1<sup>′′</sup> at 20 cm and bandwidths around half a GHz (see ). With this sensitivity and the large instantaneous field of view (FOV; around 1$`{}_{}{}^{}\times `$1 at 20 cm), SKA would detect $``$100 radio sources per square arcmin (Fig.4; see also Hopkins et al, this volume). To estimate the number of GL system, useful for particular research projects, we use the following formula: $$\mathrm{N}=\mathrm{FOV}\times n_\mathrm{r}\times r_{\mathrm{gl}}\times f_{\mathrm{r}\mathrm{type}}\times f_{\mathrm{g}\mathrm{type}},$$ (1) where FOV is the field–of–view, $`n_\mathrm{r}`$ is the number–density of radio source on the sky, $`r_{\mathrm{gl}}`$ is the strong–lensing rate of these sources, and $`f_{\mathrm{r}\mathrm{type}}`$ and $`f_{\mathrm{g}\mathrm{type}}`$ are the fractions of radio–sources and lens–galaxies with properties that one is interested in, respectively. Using either the lensing rate in the Hubble Deep Field, which has a comparable depth and number–density of objects, or the typical lensing rate for high–z sources of about $``$$`10^3`$, every synthesis observation with SKA contains a few hundred to one thousand GL systems. About 10% of these radio sources are expected to have compact structure , which will be easier to identify than the lensed starburst systems which dominate the source counts at $`\mu `$Jy levels. However, even identifying the fainter cousins of the CLASS survey, will still be a major task, with the typical distance between radio sources being $``$6<sup>′′</sup>, and not having the benefit of equally large optical images to identify the lensing galaxies. However quad–configurations and variability will be very helpful. What to do with that many GL systems? Below we discuss two projects, which we think might still be worth doing in $``$10 years time (i.e. the development and implementation time–scale for SKA), although there will undoubtedly be many more projects. Galaxy structure & evolution at high redshifts: To study the structure & evolution of galaxies up to very high redshifts, one would preferably have both their colors and mass–distribution. Colors can nowadays be obtain with instruments such as HST and ground–based 8/10–m class telescopes. However, obtaining information on the mass distribution of a significant sample of high–redshift galaxies is and will remain exceedingly difficult in the near future. This is especially true for spiral galaxies. For example, to obtain a decent sample of spiral galaxies with detailed information on their mass distribution, one would like to have Einstein–ring type GL systems (like B0218+357), which provide considerable information on the mass distribution of the lens galaxy. Taking a fraction of $``$15% for the expected fraction of spiral–lens galaxies, $``$50% for the fraction of extended radio sources (i.e. larger than the Einstein radius), $`n_\mathrm{r}`$$``$100 per sq. arcmin and $`r_{\mathrm{GL}}`$=10<sup>-3</sup>, we expect $``$30 or more spiral–lens galaxies with radio Einstein rings per FOV! Not only can these Einstein–ring structures be used to constrain the mass distributions of these spiral galaxies, they can also be used to probe the HI content of the galaxy through absorption–line measurements, simultaneously constraining their velocity fields. Especially HI–absorption against lensed extended background radio sources would be very informative (e.g. PKS1830-211; ). Constraining the mass-function of compact objects in high–redshift galaxies: As in the case of B1600+434, multi-frequency radio monitoring of those lensed sources which are compact and flat-spectrum, would allow one to separate variability resulting from strong and weak scattering by the ISM, plasma lensing and microlensing. Time delays can be determined, allowing one to determine the difference light curve between lens images. This is the only unambiguous method to separate intrinsic from non–intrinsic variability. How important this will turn out to be remains to be seen. About 10% percent of all disk–lens systems is expected to lie within 5 degrees of edge-on. In fact, the CLASS/JVAS radio–GL survey has so far discovered at least 17 new lens systems, of which B1600+434 is one. One can therefore expect $``$5% of all GL to be of this type. Assuming that $``$10% of all radio sources are compact and flat–spectrum, a few of these systems could be detected with SKA for every FOV ($``$1 sq. degree). That is, a survey with SKA could detect $``$10 of these systems per day. How could we use these systems to learn about the mass function of compact objects in their halo? The study of variability in the lens images of these systems, as a function of height above the disk of the lens galaxy, could provide unique information on the composition of their dark–matter halos. Although most lens systems would be expected to lie at redshifts between 0.3 and 1, a fraction of $``$5% will be found at redshifts $``$1.5, allowing the study of the redshift dependence of dark matter halos. Because with SKA one is looking at sources at the $`\mu `$Jy level, the expected angular size of these radio sources is $$\mathrm{\Delta }\theta \sqrt{\frac{S_{\mu \mathrm{Jy}}}{T_{12}}\left(\frac{\lambda }{25\mathrm{c}\mathrm{m}}\right)^2}\mu \mathrm{as},$$ (2) where $`T_{12}`$ is the brightness temperature of the source in units of $`10^{12}`$ K. If the source is has a redshift $`z`$ and Doppler–boosting factor $`𝒟`$, one should replace $`T_{12}`$ with $`𝒟T_{12}/(1+z)`$. Many compact flat–spectrum radio sources at high redshifts contain superluminal components, with $`𝒟`$ of order a few. For example, a superluminal 0.1 $`\mu `$Jy jet–component with $`𝒟`$$``$10 in a radio source at $`z`$=2 can be as small as 0.15 $`\mu `$as (with $`T_{12}`$$``$1), which is much smaller than the Einstein radius of a 1–M compact object at intermediate redshift! We still do not know whether the radio–faint AGN will also show superluminal motion. On the other hand, much fainter galactic sources also show superluminal motion. Such a component moving with superluminal velocity over a microlensing magnification pattern will show up in the image lightcurves as variability of up to many tens of percents on time scales of months, weeks and maybe even days. Because of the compactness of these jet–components, the variability in their light curves, induced by microlensing, will almost perfectly trace the magnification pattern due to the compact objects, as is the case in the optical (Fig.5). Moreover, for superluminal sources the microlensing time–scale is not dominated by the velocity of the compact lens objects (few hundred km/s), but by the velocity of the jet–component ($``$$`c`$). Hence, not only the variability time–scale is compressed by a factor $``$$`10^3`$, also the microlensing rate will enormously increase . Flat–spectrum radio sources could therefore be the perfect probes to study compact objects and their mass function in galaxies up to high redshifts! However, how can one seperate microlensing, scintillation and intrinsic variability of these sources? First, scintillation takes place on time-scales of $``$ few hours for these ultra–compact sources. By averaging over longer time scales one can easily remove this, leaving only microlensing and intrinsic variability. Subtracting the averaged light curves of the different lens images, as for B1600+434, will subsequently remove intrinsic variability. The difference light curve that remains only contains microlensing variability. Second, both microlensing and scintillation are strongly dependend on frequency. Whereas microlensing increases with frequency, because the source becomes more compact (eqn.2), scintillation will decrease, because of the frequency–dependent refractive properties of the Galactic ionized ISM (e.g. Fig.3). The scintillation is of course strongest around the transition frequency from weak to strong, about 5 GHz. At longer wavelengths, e.g. 21cm, very fast diffractive scintillation could also become important for sub–microarcsecond sources. The large bandwith of SKA ($`\mathrm{\Delta }\nu `$$``$$`\nu /2`$), however, will allow one to further discriminate between microlensing and the different types of scintillation. In the case of B1600+434, separating properties of the source and mass–function is complicated. In the case of very compact sources, which will be observed with SKA, this is less problematic. As mentioned before, these sources are usually much smaller than the angular scale over which the magnification pattern changes significantly. In other words, one can often regard the source as a point source, which clearly simplifies unraveling the source and mass–function properties. The statistical properties of the difference light curves can thus be used to study the mass function of compact objects at those positions where the lens images pass through the lens galaxy (e.g. halo, disk or bulge). Microlensing studies, of course, would not have to be confined to multiply–imaged radio sources. Detailed multi–frequency studies of suitable alignments of compact radio sources and foreground galaxies could be used to study the properties of lensing matter as a function of distance to the galaxy out to any distance, far into its dark–matter halo! A statistical study of the frequency–dependent modulation characteristics as a function of distance to the foreground galaxy can also be used to separate galactic scintillation from microlensing because the galactic ISM is not expected to change rapidly on angular scales of a few arcseconds. ## 6 Conclusions Illustrated by the GL system B1600+434, we have shown the current limitations, but also the exciting possibilities of detecting compact objects in the dark–matter halo of intermediate–redshift galaxies and contraining their mass function through radio microlensing. The Square Kilometer Array will be two orders of magnitude more sensitive than the current VLA, and, because of multi–beaming, an additional order of magnitude. Not only can SKA find $``$10 GL systems comparable to B1600+434 per day using its higher resolution, larger bandwidth and continuous monitoring capabilities (because of the multi–beam setup), we have also shown that microlensing, scintillation and intrinsic variability can easily be separated in these systems. SKA will therefore be a unique instrument to study compact objects and their mass–function in galaxies and their dark–matter halo up to very high redshifts. The large increase in the microlensing rate (a factor $``$1000 compared with stationary optical sources) for lensed flat–spectrum radio sources with superluminal components makes this a nearly impossible task for any optical telescope. ## References
warning/0001/cond-mat0001291.html
ar5iv
text
# Coexistence of excited states in confined Ising systems ## Abstract Using the density-matrix renormalization-group method we study the two-dimensional Ising model in strip geometry. This renormalization scheme enables us to consider the system up to the size $`300\times \mathrm{}`$ and study the influence of the bulk magnetic field on the system at full range of temperature. We have found out the crossover in the behavior of the correlation length on the line of coexistence of the excited states. A detailed study of scaling of this line is performed. Our numerical results support and specify previous conclusions by Abraham, Parry, and Upton based on the related bubble model. The understanding of classical systems in confined geometries has been a challenge for several years . Among such investigated systems are fluids or magnets confined between parallel walls. Studies of finite-size effects have not been limited only to the vicinity of the critical point, but also to the first-order phase transitions, which are less known. In this Raport we consider the two-dimensional Ising system on a square lattice in strip geometry ($`L`$ is width of the strip) with the Hamiltonian $$=J\underset{<i,j>}{}\sigma _i\sigma _jH\underset{i}{}\sigma _i,$$ (1) where the coupling $`J>0`$, $`H`$ is the bulk magnetic field and $`\sigma _i=\pm 1`$. The first sum runs over all nearest-neighbour pairs of sites while the second sum runs over all sites. Even such a simple model has an interesting crossover governed by the bulk magnetic field , which value $`H_x`$ depends on temperature and the size of the system. The borderline $`H_x(T;L)`$ divides two different $`L`$ and $`H`$ dependencies of the correlation length $`\xi `$. Using the bubble model Abraham et al. found that at subcritical temperatures one has $`1/\xi =P(T)L|H|,\mathrm{for}\mathrm{\hspace{0.17em}\hspace{0.17em}\hspace{0.17em}0}<|H|H_x,`$ (2) $`1/\xi =R(T)+S(T)|H|^{2/3},\mathrm{for}|H|H_x,`$ (3) where $`P(T)=2m/k_BT`$, $`R(T)=2\sigma _0/k_BT`$. Here, $`S(T)`$ is an unknown positive coefficient. Furthermore, $`m`$ and $`\sigma _0`$ are the bulk spontaneous magnetization and the interfacial tension, respectively. The bubble model studies concluded that $`H_x(T;L)`$ scales towards the first-order line according to the form : $$H_x(L;T)A(T)L^\alpha +B(T)L^\gamma +C(T)L^\delta +\mathrm{},$$ (4) where $`\alpha =1`$, $`\gamma =5/3`$, and $`\delta =7/3`$. A similar problem of higher-order corrections, but to the Kelvin equation (the scaling of the bulk coexistence field in the presence of the parallel surface fields) has been studied recently . Using the density-matrix renormalization-group method (DMRG) it was found that for a large range of surface fields and temperatures corrections are not compatible with the behavior predicted by the existing theory . It is one of reasons why we have checked out here the predictions given by the bubble model. Abraham et al. argued that the mentioned crossover occurs because the class of dominating configurations determining the behavior of correlation functions changes from a single connected loop for $`|H|>H_x`$ to two disconnected closed loops $`|H|<H_x`$ (in cylinder geometry). In our case, where the free boundaries are present, for $`|H|<H_x`$, the dominating configurations consist of succeeding pieces of a strip with opposite magnetizations . For $`|H|>H_x`$ the most important configurations contributing to the correlation function are again closed loops including domains of opposite magnetizations (see, Fig.1). In order to analyze this problem beyond the bubble model, we can use the transfer-matrix (TM) calculations . However, it is well known that to obtain satisfactory finite-size scaling results, one should consider large enough systems . This may, in turn, complicate calculations or even make them impossible. To overcome this problem we have applied the DMRG method for two-dimensional systems based on the TM approach. Providing a very efficient algorithm for the construction of the effective transfer matrices for large $`L`$ this method was successfully employed for a number of problems (for which no exact solutions are available, e.g. for nonvanishing bulk fields) . Using it we were able to analyze the system in full range of temperatures and the bulk magnetic field for strips of widths up to $`L=300`$. For a comprehensive review of background, achievements and limitations of DMRG, see Ref. . We first calculated the free-energy levels $$f_i(H,T;L)=\frac{k_BT}{L}\mathrm{ln}\left(\lambda _i(H,T;L)\right),$$ (5) for $`i=0,1,2,\mathrm{}`$, where $`\lambda _i`$ are the eigenvalues of the TM arranged in order of decreasing magnitude. Because the inverse (longitudinal) correlation length can be defined as $$1/\xi (L)=\mathrm{log}(\lambda _0/\lambda _1),$$ (6) and the lowest free-energy level does not cross others, especially important are the values of the bulk magnetic field $`H_x(T;L)`$, where the first and the second excited states cross each other. In such a case we can observe the crossover in the behavior of the correlation length. Let us first analyze the structure of the TM low-lying levels as a function of the bulk magnetic field $`H`$ at fixed $`T`$. At very low temperature they should behave practically in the same way as the ground state energy. Therefore, it is worthy first considering the ground state properties of the system. Let us define the configuration of a row for the strip in the following way $`|\sigma _1,\sigma _2,\mathrm{},\sigma _{L1},\sigma _L`$, where the values of $`\sigma _i`$ are denoted $`\pm `$ for simplicity. At zero magnetic field $`H`$ the two states with all spins positive $`|++\mathrm{}++`$ or negative $`|\mathrm{}`$ have the same energy. The extra magnetic field term splits both states and the energy per spin is $$ϵ_{1,2}=J(2\frac{1}{L})\pm H.$$ (7) Assuming $`H>0`$ the $`|++\mathrm{}++`$ state is always the singlet ground state. In order to find the first excited states we have to flip the first or the last column ($`i=1,L`$) in the previous configurations. In this way we get the four states $`|+\mathrm{}++`$, $`|++\mathrm{}+`$, $`|+\mathrm{}`$ and $`|\mathrm{}+`$. The magnetic field splits this level into two doublets and for the two first states their energy decreases when the $`H`$ increases according to the equation $$ϵ_{3,4}=J(2\frac{3}{L})H(1\frac{2}{L}).$$ (8) Therefore, we expect the crossing of the singlet state $`|\mathrm{}`$ with the doublet $`|+\mathrm{}++`$, $`|++\mathrm{}+`$ at a certain value of the bulk magnetic field $$H_x(T=0;L)=\frac{J}{L1}.$$ (9) Note, that for $`T0`$ Eq.(4) reduces to Eq.(9) provided $`A(T)J`$ and $`B(T),C(T)0`$. At finite temperatures we do not have any more real crossing points but, so-called “the regions of avoided level crossing”. At $`H=0`$ the first two levels are separated as $`f_1f_0\mathrm{exp}(\sigma _0L/k_BT)`$, so they are asymptotically degenerated for $`L\mathrm{}`$. The region of avoided level crossing continues for nonzero magnetic fields up to $`|H|\mathrm{exp}(\sigma _0L/k_BT)`$. When we are interested in the behavior of $`\xi `$, we have to consider the second and third eigenvalues of the transfer matrix , where asymptotic degeneracy is also present for $`f_1`$ and $`f_2`$ . It is assumed that the difference $`f_2f_1`$ and the avoided level crossing region centered on $`H_x`$ are of order $`\mathrm{exp}(CL)`$, where the coefficient C may be $`H`$ and $`T`$ dependent. It makes sense to discuss algebraic shift of the value $`H_x(T;L)`$ for $`L\mathrm{}`$. In order to find out the value of $`H_x`$ in finite temperatures at fixed $`L`$, we identify $`H_x`$ with the value of the bulk magnetic field where the second free-energy level $`f_1`$ has a maximum and the separation $`f_2f_1`$ is the smalest one. The curves $`H_x(T;L)`$ present the coexistence of the excited states and are shown in Fig.2. The curves indicate the phase boundaries between the two phases with different dependencies of $`\xi `$ on $`L`$ and $`H`$. As $`L\mathrm{}`$ the coexistence lines of the excited states shift towards the $`H=0`$ axis that is intuitively clear at $`T=0`$. Since the width of the strip increases the energy of configurations where only one column of spins is flipped (Eq.(8)) are close and close to the energy of configurations with all spins pointed in one direction (Eq.(7)), so in the $`L\mathrm{}`$ limit one has $`H_x=0`$. Let us discuss the scaling of the coexistence line $`H_x(T;L)`$ to the bulk first-order line (Fig.(2)). To verify the bubble model predictions (Eq.(4)) we have calculated series of values of $`H_x(T=\mathrm{const};L)`$ for $`L=20,40,\mathrm{},200`$ and for temperatures ranging from $`T0.44T_c`$ up to $`T0.99T_c`$. Table. Scaling exponents of $`H_x(T;L)`$ extrapolated from the DMRG data. $`T`$ $`\alpha `$ $`\gamma `$+1 $`\delta `$+5/3 1.00 -0.9994(5) -0.668(1) -0.66(4) 1.50 -0.9990(5) -0.667(1) -0.64(1) 1.75 -1.000(1) -0.668(2) -0.64(1) 2.00 -0.998(1) -0.667(1) -0.67(1) 2.15 -1.028(3) -0.67(1) -0.67(2) 2.20 -1.002(6) -0.69(1) -0.68(3) Table shows the values of scaling exponents obtained from the DMRG data. Using the powerful extrapolation technique, the Bulirsch and Stoer (BST) method , we obtained an excellent agreement with Abraham et al.. In order to get the $`A`$ coefficient in Eq.(4) one can compare Eqs. (2) and (3). They have to agree at the value $`H=H_x`$ in the termodynamic limit, which implies the following relation : $$A(T)=\sigma _0(T)/m(T).$$ (10) In Fig.3 our data reconstruct this curve very well. To the best of our knowledge, the coefficients $`B(T)`$ and $`C(T)`$ in Eq.(4) have not been yet determinated, but our numerical results can predict their temperature behavior. Close to $`T_c`$ the validity of Eq.(4) is limited because the scaling of points of the $`H_x`$ curve is governed by the bulk critical point. In order to study it in detail we have considered characteristic points of the upper part of the $`H_x`$ curve: the inflection points ($`H_c(L)`$,$`T_c(L)`$) and the end points $`T^{}(L)`$ (see, Fig.(2)), where the following scaling is expected: $`\tau _c(L)=(T_\mathrm{c}T_\mathrm{c}(L))/T_\mathrm{c}L^{y_T},`$ (11) $`H_\mathrm{c}(L)L^{y_H}.`$ (12) Here, $`y_T=1`$ and $`y_H=15/8`$ are the thermal and magnetic exponents of the two-dimensional Ising model. To verify the scaling to the critical point ($`H=0`$,$`\tau =0`$) we found out the inflection points and the end points for $`L=30,60,100,130,160`$ and $`200`$ using subsequently the BST technique. We have examined the scaling form (12) for $`L\mathrm{}`$ and found very good agreement $`\tau _c=0.00006(6),\mathrm{and}y_T=1.005(5)`$ (13) $`H_c=0.0006(6),\mathrm{and}y_H=1.876(8)`$ (14) $`\tau ^{}=0.0000(3),\mathrm{and}y_T=1.006(6)`$ (15) Note, that for $`TT_c`$ and $`L\mathrm{}`$ we can reproduce the scaling form (12) from Eq.(4) by assuming that $`A(T)0`$, $`B(T)0`$, and $`C(T)\mathrm{}`$. This is in agreement with our numerical estimations for scaling coefficients as depicted in Fig.3. Of course, this relation is not valid at $`T_c`$. In order to analyze the behavior of the correlation length we have derived $`1/\xi `$ for $`L`$ between $`100`$ and $`300`$ in temperatures below $`T_c(L)`$. To examine the form of Eq.(2) firstly we have confirmed the linear dependence of the coefficient on $`L`$. Next we compared our numerical results with the coefficients $`P(T)`$ and $`R(T)`$ in Eqs.(2,3). What is more, we presented the temperature dependence of the $`S(T)`$ coefficient which was not determined in the bubble model (see, Fig.4). When temperature raises, more and more complex configurations on the Ising strip contribute to the free energy in contrast to the assumption of the bubble model . Consequently, in high temperatures the validity of Eqs. (2,3) is limited to a narrow range of $`H`$. The bubble model predictions are also spoiled by the presence of strong bulk magnetic field. That is why, the higher is temperature the smaller $`H`$ is necessary to recover the linear dependence of $`1/\xi `$ on $`H`$, as in Eq.(2). Similarly, when $`TT_c(L)`$ the regime with the $`H^{2/3}`$ dependence of $`1/\xi `$ (Eq.(3)), close to the right side of the coexistence line, shrinks to zero. In conclusion, we have used the density-matrix renomalization-group method to obtain reliable information about the two-dimensional Ising model in the bulk magnetic field. We have confirmed the crossover related to the correlation length analyzed before for the bubble model . Our study has not been limited to subcritical temperatures and small bulk fields. We have confirmed Abraham et al. predictions for the scaling of the first-order line in the subcritical region. Morover, we have established the precise scaling form for the bulk magnetic field by numerically determining coefficients $`B(T)`$ and $`C(T)`$ in Eq.(4). Furthermore, we have extended the analysis of the bubble model to critical region verifying that the scaling behavior is governed by the bulk critical point. Finally, we numerically confirmed the magnetic field dependence of the correlation length, simultaneosly extracting previously unknown coefficient $`S(T)`$ in Eq.(3). Above results demonstrate that for two-dimensional classical systems the DMRG technique provides significantly accurate data for studying equilibrium properties of large systems in a nonvanishing bulk magnetic field. ###### Acknowledgements. I thank A.O. Parry for suggesting me the topic of this Raport. I am grateful to E. Carlon and T.K. Kopeć for critical reading of the manuscript. This work was supported by the Polish Science Committee (KBN) under Grant No. 2P03B10616.
warning/0001/math0001101.html
ar5iv
text
# Integrality of 𝐿²-Betti numbers ## 1 Introduction ###### 1.1 Remark. This is a corrected version of an older paper with the same title. The proof of one of the basic results of the earlier version contains a gap, as was kindly pointed out to me by Pere Ara. This gap could not be fixed. Consequently, in this new version everything based on this result had to be removed. In Atiyah introduced $`L^2`$-Betti numbers of closed manifolds in terms of the kernel of the Laplacian on the universal covering. There, he asked what the possible values of these numbers are, in particular whether they are always integers if the fundamental group is torsion-free. We call this the Atiyah conjecture. More precisely, we consider the following algebraic question: ###### 1.2 Definition. Let $`\mathrm{\Lambda }`$ be an additive subgroup of the rationals, and $`K`$ a subring of $``$. We say a discrete group $`G`$ fulfills the *Atiyah conjecture of order $`\mathrm{\Lambda }`$ over $`KG`$* if $$dim_G(\mathrm{ker}A)\mathrm{\Lambda }AM(m\times n,KG),$$ where $`\mathrm{ker}A`$ is the kernel of the induced map $`A:l^2(G)^nl^2(G)^m`$. Let $`\mathrm{pr}_{\mathrm{ker}A}`$ be the projection onto $`\mathrm{ker}A`$. Then $$dim_G(\mathrm{ker}A):=\mathrm{tr}_G(\mathrm{pr}_{\mathrm{ker}A}):=\underset{i=1}{\overset{n}{}}\mathrm{pr}_{\mathrm{ker}A}e_i,e_i_{l^2(G)^n},$$ where $`e_il^2(G)^n`$ is the vector with the trivial element of $`Gl^2(G)`$ at the $`i^{\text{th}}`$-position and zeros elsewhere. $`\mathrm{tr}_G`$ is the canonical finite trace on the von Neumann algebra of operators on $`l^2(G)^n`$ commuting with the right $`G`$-action, the so called *Hilbert $`𝒩G`$-module maps*. The group $`G`$ is said to fulfill the *strong Atiyah conjecture over $`KG`$* if $`\mathrm{\Lambda }`$ is generated by $`\{\left|F\right|^1|F<G\text{ finite}\}`$. ###### 1.3 Remark. In it is shown that the so called Lamplighter group $`G`$ does not satisfy the Atiyah conjecture over $`G`$. However, there is no bound on the orders of finite subgroups of $`G`$, in contrast to all examples of groups for which the strong Atiyah conjecture is known. In Definition 1.2, we can replace the kernels by the closures of the images or by the cokernels because these numbers by additivity of the $`G`$-dimension \[20, 1.4\] pairwise sum up to the dimension of the domain or range, which is an integer. Replacing $`A`$ by $`A^{}A`$ we may also assume that $`n=m`$. Because we can multiply $`A`$ with a unit in $`Q`$ without changing its kernel, the Atiyah conjecture over $`KG`$ implies the Atiyah conjecture over $`QG`$, where $`Q`$ is the quotient field of $`K`$ in $``$. Usually therefore we will assume that $`K`$ is a field. The Atiyah conjecture of order $`\mathrm{\Lambda }`$ over $`G`$ is equivalent to the statement that all $`L^2`$-Betti numbers of finite $`CW`$-complexes with fundamental group $`G`$ are elements of $`\mathrm{\Lambda }`$ \[20, Lemma 2.2\], i.e. to the original question of Atiyah. Fix a coefficient field $`K`$. If not stated otherwise, the term “Atiyah conjecture” means “Atiyah conjecture over $`KG`$”. We introduce a new method of proof for the Atiyah conjecture which is based on the approximation results in . In it is proved that in many cases if a group $`G`$ is the direct or inverse limit of a directed system of groups $`G_i`$, then $`L^2`$-Betti numbers over $`G`$ are limits of $`L^2`$-Betti numbers defined over the groups $`G_i`$. More precisely: ###### 1.4 Definition. (compare \[26, 1.11\]) Let $`𝒢`$ be the smallest class of groups which contains the trivial group, is subgroup closed and is closed under direct or inverse limits of directed systems, and under extensions with amenable quotient. The amenable quotient is not necessarily a quotient group, one relaxes this to a suitable notion of an amenable (discrete) quotient space (by a subgroup which is not normal) (compare \[26, 4.1\]). This class of groups is very large. It contains most groups which naturally occur in geometry, e.g. all residually finite groups. It would be interesting to find an example of a group which does not belong to $`𝒢`$. ###### 1.5 Proposition. Let $`G_i𝒢`$ (with $`𝒢`$ as in Definition 1.4) be a directed system of torsion-free groups which fulfill the strong Atiyah conjecture. Then their (direct or inverse) limit $`G`$ fulfills the strong Atiyah conjecture over $`G`$. ###### 1.6 Corollary. The strong Atiyah conjecture over $`G`$ is true if $`G`$ is a pure braid group, or a positive $`1`$-relator group (in the sense of ) with torsion free abelianization. The corollary will be explained in Example 2.8. There is one significant difference to all other results about the Atiyah conjecture obtained so far: the approximation result is proved only over $`G`$ instead of $`G`$. This does not effect the original question of Atiyah. But it is relevant for the zero divisor conjecture (by Lemma 1.7). In a forthcoming paper we will show how to enlarge $``$ at least to the field of algebraic numbers in $``$. Proposition 1.5 can be used to give an alternative proof of the Atiyah conjecture for free groups, compare Example 2.5. Linnell obtained the most general positive results about the Atiyah conjecture so far . His interest stems from the zero divisor conjecture and the following observation (compare e.g. ): ###### 1.7 Lemma. If for a torsion-free group the strong Atiyah conjecture is true over $`KG`$ then the ring $`KG`$ has no non-trivial zero divisors. Linnell uses essentially two lines of argument for the Atiyah conjecture: algebraic considerations using $`K`$-theoretic information (in particular Moody’s induction theorem \[23, Theorem 1\]) to obtain the Atiyah conjecture for elementary amenable groups or for extensions with elementary amenable quotient. For a free group $`F`$, Linnell devised a completely different argument, making use of Fredholm module techniques which were used before to prove that there are no non-trivial projectors in $`C_r^{}F`$. Linnell defines : ###### 1.8 Definition. The class $`𝒞`$ is the smallest class of groups which contains all free groups and is closed under directed union and extensions with virtually abelian quotients. The class $`𝒞^{}`$ consists of extensions of direct sums of free groups with elementary amenable quotient. Putting his two approaches together, Linnell proves in \[17, 1.5\] ###### 1.9 Theorem. The strong Atiyah conjecture over $`G`$ is true for groups in the classes $`𝒞`$ which have a bound on the orders of finite subgroups. The proof of the Atiyah conjecture for the class $`𝒞^{}`$ given in contains a gap (which parallels the gap mentioned in Remark 1.1). However, we will shown in Corollary 4.4 how to fill this gap and conclude that the strong Atiyah conjecture over $`G`$ is true if $`G𝒞^{}`$. We use the new results to define the following class of groups for which the Atiyah conjecture is true: ###### 1.10 Definition. Let $`𝒟`$ be the smallest non-empty class of groups such that: 1. If $`G`$ is torsion-free and $`A`$ is elementary amenable, and we have a projection $`p:GA`$ such that $`p^1(E)𝒟`$ for every finite subgroup $`E`$ of $`A`$, then $`G𝒟`$. 2. $`𝒟`$ is subgroup closed. 3. Let $`G_i𝒟`$ be a directed system of groups and $`G`$ its (direct or inverse) limit. Then $`G𝒟`$. Observe that $`𝒟`$ contains only torsion-free groups. ###### 1.11 Theorem. If $`G𝒟`$ then the strong Atiyah conjecture over $`G`$ is true for $`G`$. ###### 1.12 Corollary. If $`G𝒟`$ then the ring $`G`$ does not contain non-trivial zero divisors. ###### 1.13 Proposition. All torsion-free groups in $`𝒞`$ and $`𝒞^{}`$ are also contained in $`𝒟`$. Moreover, $`𝒟`$ is closed under direct sums, direct products and free products. $`𝒞`$ is not closed under direct sums. $`𝒞^{}`$ has this property, but it is not closed under free product. $`𝒟`$ is an enlargement which repairs these deficits. ###### 1.14 Example. It follows that $`G:=(\times )`$ fulfills the Atiyah conjecture over $`G`$. The $`L^2`$-Betti numbers of $`G`$ are computed as $`b_1^{(2)}(G)=1=b_2^{(2)}(G)`$ and $`b_k^{(2)}(G)=0`$ for $`k1,2`$. By \[24, 9.3\] $`G𝒞`$. Using ideas of Reich, one can also show $`G𝒞^{}`$, so that indeed we cover additional groups. ### Organization of the paper Section 2 gives an overview over elementary results about the Atiyah conjecture. We also prove Proposition 1.5 and discuss applications of this proposition. In Sections 3 and 4 we prove Theorem 1.11 and Proposition 1.13. Section 5 describes possible generalizations. In Section 6 we extend Linnell’s proof of the Atiyah conjecture for free groups to another class of groups, the so called treelike groups, and characterize the treelike groups. ## 2 The Atiyah conjecture for limits of groups Here we collect a few statements where the Atiyah conjecture for some groups implies its validity for other groups. The following propositions are well known. ###### 2.1 Proposition. (\[18, 8.6\]) If $`H`$ is a subgroup of index $`n`$ in $`G`$ and $`H`$ fulfills the Atiyah conjecture of order $`\mathrm{\Lambda }`$ over $`KH`$, then $`G`$ fulfills the Atiyah conjecture of order $`\frac{1}{n}\mathrm{\Lambda }`$ over $`KG`$. ###### 2.2 Proposition. If $`G`$ fulfills the Atiyah conjecture of order $`\mathrm{\Lambda }`$ over $`KG`$, and if $`U`$ is a subgroup of $`G`$, then $`U`$ fulfills the Atiyah conjecture of order $`\mathrm{\Lambda }`$ over $`KH`$. ###### Proof. This follows from the fact that the $`U`$-dimension of the kernel of a matrix over $`KU`$ acting on $`l^2(U)^n`$ coincides with the $`G`$-dimension of the same matrix, considered as an operator on $`l^2(G)^n`$ \[26, 3.1\]. ∎ ###### 2.3 Proposition. Let $`G`$ be the directed union of groups $`\{G_i\}_{iI}`$ and assume that each $`G_i`$ fulfills the Atiyah conjecture of order $`\mathrm{\Lambda }_i`$ over $`KG_i`$. Then $`G`$ fulfills the Atiyah conjecture of order $`\mathrm{\Lambda }`$ over $`KG`$, where $`\mathrm{\Lambda }`$ is the additive subgroup of $``$ generated by $`\{\mathrm{\Lambda }_i\}_{iI}`$. ###### Proof. A matrix over $`KG`$, having only finitely many non-trivial coefficients, already is a matrix over $`KG_i`$ for some $`i`$. The $`G_i`$-dimension and the $`G`$-dimension of the kernel of the matrix coincide by \[26, 3.1\]. ∎ For groups in $`𝒢`$ we have good approximation results for $`L^2`$-Betti numbers \[26, 6.9\]. We use these to give a proof of the Atiyah conjecture for suitable groups. For this it is necessary to work over the rational group ring instead of the complex group ring. ###### 2.4 Proposition. Let $`G_i`$ be a directed system of groups which fulfill the Atiyah conjecture over $`G_i`$ of common order $`\frac{1}{L}`$. Suppose all the groups $`G_i`$ lie in the class $`𝒢`$ (defined in Definition 1.4). Then their direct or inverse limit $`G`$ fulfills the Atiyah conjecture of order $`\frac{1}{L}`$ over $`G`$. This is a direct consequence of the assumption and the approximation result \[26, 6.9\], since there it is shown that each $`L^2`$-Betti number over the limit group is the limit of $`L^2`$-Betti numbers over the groups in the sequence. By assumption, all of these lie in $`\frac{1}{L}`$, which is a closed subset of $``$. Therefore the same is true for the limit $`L^2`$-Betti number. Note that this implies the strong Atiyah conjecture if we know that $`L`$ is the least common multiple of the orders of finite subgroups of $`G`$ (provided this number exists), in particular if all $`G_i`$ are torsion-free and fulfill the strong Atiyah conjecture. Hence Proposition 1.5 is a special case of Proposition 2.4. ###### 2.5 Example. We use Proposition 2.4 to give a different proof that a free group $`F`$ fulfills the strong Atiyah conjecture over $`F`$. By free groups are residually torsion-free nilpotent. More precisely, if we consider the descending central series $`FF_1F_2\mathrm{}`$ then $`F_k=\{1\}`$ and $`F/F_k`$ is torsion-free. It follows that the groups $`F/F_k`$ are nilpotent and poly-$``$. Using the methods of the ring-theoretic parts of it is easy to prove the strong Atiyah conjecture (over the complex group ring) for poly-$``$ groups. We will repeat the proof in Section 3. Alternatively, one can rely on a different type of ring theory and in particular avoid the use of the rings $`DG`$ and $`𝒰G`$ of Definition 3.3: If $`G`$ is poly-$``$, then $`G`$ is Noetherian \[25, 8.2.2\]. Using Moody’s induction theorem \[23, Theorem 1\] we see that the $`G`$-theory of $`G`$ is trivial, which by \[28, 4.6\] coincides with the $`K`$-theory. Therefore a finite projective resolution of a $`G`$-module can be replaced by a finite free resolution. By \[25, 8.2.18\] $`G`$ has finite cohomological dimension. Together this implies that every finitely generated module over $`G`$ has a finite free resolution. If $`A`$ is a matrix over $`G`$, its kernel in $`(G)^n`$ is finitely generated and we define its $`G`$-dimension as the alternating sum of the ranks of the modules in a finite free resolution. One can prove that this number, which is an integer, coincides with the $`G`$-dimension of $`\mathrm{ker}Al^2(G)^n`$ (this is e.g. done in the thesis \[7, 5.4.1\]). We therefore arrive at the inverse system $`\mathrm{}F/F_2F/F_1`$ where each of the quotients $`F/F_k`$ is torsion-free and fulfills the strong Atiyah conjecture and lies in $`𝒢`$. $`F`$ is a subgroup of the inverse limit of this system. Application of Proposition 1.5 and Lemma 2.2 gives the result. There is of course one drawback: the approximation method proves the Atiyah conjecture only for matrices over $`F`$ instead of $`F`$. In view of Proposition 2.4 the following lemma is of importance. ###### 2.6 Lemma. The class $`𝒟`$ is contained in $`𝒢`$. ###### Proof. By definition, $`𝒢`$ is subgroup closed and closed under direct and inverse limits of directed systems, and under extensions with elementary amenable quotients. It follows that $`𝒟`$ is contained in $`𝒢`$. ###### 2.7 Corollary. The strong Atiyah conjecture over $`G`$ is true if $`G`$ is residually torsion free of class $`𝒞`$, in particular if $`G`$ is residually torsion-free solvable (or more generally residually torsion-free elementary amenable). ###### Proof. Let $`𝒳`$ be a class of groups. We call a group $`G`$ residually of class $`𝒳`$ if $`G`$ has a sequence of normal subgroups $`GH_1H_2\mathrm{}`$ with $`_kH_k=\{1\}`$ and such that the quotients $`G/H_k𝒳`$ $`k`$. Another (more common) definition of residuality is to require for every $`gG`$ a homomorphism $`\varphi _g:GX_g`$ with $`\varphi _g(g)1`$ where $`X_g𝒳`$. If $`𝒳`$ is closed under direct sums then for countable groups $`G`$ the two definitions are equivalent. This is the case for the class of torsion-free elementary amenable groups (but not for the class $`𝒞`$). If $`G`$ has a sequence of normal subgroups $`H_1H_2\mathrm{}`$ with trivial intersection such that $`G/H_k𝒞`$ and $`G/H_k`$ are torsion-free, then $`G`$ is a subgroup of the inverse limit of $`(G/H_k)_k`$. Combining Lemma 2.6 and Proposition 1.13, $`𝒞𝒢`$. Because of Theorem 1.9 we can apply Proposition 1.5 and Lemma 2.2 to conclude that $`G`$ fulfills the strong Atiyah conjecture over $`G`$. ∎ ###### 2.8 Example. Let $`F`$ be a free group on the free generators $`\{a_i\}_i`$ and assume $`cF`$ generates its own centralizer. Choose $`V`$. Let $`A`$ be a free abelian countably generated group $`A`$ and $`r`$ one of the free generators. The following groups are residually torsion free of class $`𝒞`$ and hence by Corollary 2.7 fulfill the strong Atiyah conjecture over the rational group ring: 1. Pure (also called colored) braid groups. 2. Positive $`1`$-relator groups $`G`$ with $`H_1(G,)`$ torsion free. A group is a $`1`$-relator group if it has a presentation $$G=<g_1,\mathrm{},g_n|R:=\underset{k=1}{\overset{N}{}}g_{i_k}^{n_k}=1>$$ with one relation $`R`$. It is called positive if one finds such a presentation with $`n_k0`$ for all $`k`$. By \[21, IV.5.2.\] $`G`$ is torsion free if and only if $`R`$ is not a proper power in the free group generated by $`g_1,\mathrm{},g_n`$. It follows from the presentation that $`H_1(G)=G/[G,G]`$ is torsion free if and only if the greatest common divisor of $`s_1,\mathrm{},s_n`$ is $`1`$, where $`s_r=_{i_k=r}n_k`$ is the exponent sum of $`g_r`$ in $`R`$. Observe that therefore $`H_1(G,)`$ torsion free for a positive $`1`$-relator group implies that $`G`$ is torsion free. 3. $`E_1:=<r,a_i;iV|c=r^n>`$; 4. $`E_2:=<a_i;i=1,\mathrm{},k|a_1^{n_1}a_2^{n_2}\mathrm{}a_k^{n_k}=1>`$, with $`k>3`$ and $`n_i`$; 5. $`E_3:=F_{c=r}A`$. ###### Proof. First, we deal with the positive $`1`$-relator groups. If $`R=1`$, $`G`$ is free and we have nothing to prove. By \[5, Theorem 1\] positive $`1`$-relator groups are residually solvable. We want to prove that all the derived series quotients $`G/G^{(n)}`$ are torsion free. We can assume that $`R1`$, and since all $`n_k`$ are non-negative, the second differential in the presentation $`2`$-complex $`X^{(2)}`$ of $`G`$ is non-trivial (i.e. at least one of the $`s_k`$ is not zero), and therefore $`H_2(X^{(2)},)=0`$. By assumption $`H_1(G,)`$ is torsion free. Now \[27, Theorem A\] implies that $`G/G^{(n)}`$ is torsion free for all $`n`$. By \[14, Theorem 2.6\], the pure braid groups are residually torsion-free nilpotent. The group $`E_1`$ is residually torsion-free nilpotent by \[4, Theorem 1\]. $`E_2`$ is residually free by \[3, p. 414\]. $`E_3`$ is residually free \[3, Theorem 8\]. Free groups and therefore $`E_1,E_2,E_3`$ are residually torsion-free solvable. Every nilpotent and every solvable group is contained in $`𝒞`$. ∎ ###### 2.9 Remark. In a condition is given when the Atiyah conjecture for a group $`H`$ implies the Atiyah conjecture for finite extensions of $`H`$. The pure braid groups satisfy this condition, so that in particular the strong Atiyah conjecture for the full braid groups follows. The condition is also satisfied for many positive one-relator groups. ## 3 Extensions with elementary amenable quotient The following proposition is implicit in Linnell’s work. He only deals with the complex group ring, therefore we give a complete proof here. ###### 3.1 Proposition. Let $`1HGA1`$ be an exact sequence of groups. Assume that $`G`$ is torsion free and $`A`$ is elementary amenable. For every finite subgroup $`E<A`$ let $`H_E`$ be the inverse image of $`E`$ in $`G`$. Assume that $`K=\overline{K}`$ and for all finite subgroups $`E<G`$ that $`H_E`$ fulfills the strong Atiyah conjecture over $`KH_E`$. Then $`G`$ fulfills the strong Atiyah conjecture over $`KG`$. ###### 3.2 Corollary. Suppose $`H`$ is torsion-free and fulfills the strong Atiyah conjecture over $`KH`$ with $`K=\overline{K}`$. If $`G`$ is an extension of $`H`$ with elementary amenable torsion-free quotient then $`G`$ fulfills the strong Atiyah conjecture. ###### Proof. By assumption, the only finite subgroup of $`G/H`$ is the trivial group and the Atiyah conjecture is true for its inverse image $`H`$. ∎ We essentially follow Linnell’s lines in to prove Proposition 3.1. First we repeat a few lemmas. ###### 3.3 Definition. Given the group von Neumann algebra $`𝒩G`$ we define $`𝒰G`$ to be the algebra of all unbounded operators affiliated to $`𝒩G`$ (compare e.g. \[18, Section 8\]), i.e. such that all their spectral projections belong to $`𝒩G`$. We have to consider $`KG𝒩G𝒰G`$. Let $`DG`$ be the division closure of $`KG`$ in $`𝒰G`$, i.e. the smallest subring of $`𝒰G`$ which contains $`KG`$ and which has the property that whenever $`xDG`$ is invertible in $`𝒰G`$ then $`x^1DG`$. The Atiyah conjecture is strongly connected to ring theoretic properties of $`DG`$: ###### 3.4 Lemma. Let $`G`$ be a torsion-free group. $`G`$ fulfills the strong Atiyah conjecture over $`KG`$ if and only if the division closure $`DG`$ of $`KG`$ in $`𝒰G`$ is a skew field. ###### Proof. The “only if” part is proved in . Linnell states the result only for $`K=`$ but the proof carries over verbatim to the more general case. For the converse suppose $`DG`$ is a skew field. We have to show that for $`AM(m\times n,KG)`$ the $`G`$-dimension of $`\mathrm{coker}(A_{KG}\mathrm{id}_{l^2G})`$ is an integer, or equivalently that the $`G`$-dimension of $`\mathrm{coker}(A_{KG}\mathrm{id}_{𝒰G})`$ is an integer. Since tensor products are right exact, this amounts to the fact that the $`G`$-dimension of $`(\mathrm{coker}(A)_{KG}DG)_{DG}𝒰G`$ is integral. However, $`\mathrm{coker}(A)_{KG}DG`$ is a (finitely generated, since $`A`$ is a finite matrix) module over the skew field $`DG`$, therefore isomorphic to $`(DG)^N`$ for a suitable integer $`N`$. Then $`\mathrm{coker}(A)_{KG}𝒰G(𝒰G)^N`$ which has $`G`$-dimension $`N`$. ∎ ###### 3.5 Lemma. Suppose $`H`$ is a normal subgroup of $`G`$ with infinite cyclic quotient. Let $`(𝒰H)G`$ be the subring of $`𝒰G`$ generated by $`G`$ and $`𝒰H`$. If $`tGH`$ and $`x=1+q_1t+\mathrm{}+q_kt^k(𝒰H)G`$ with $`q_i𝒰H`$ then $`x`$ is invertible in $`𝒰G`$ and in particular is not a zero divisor. ###### Proof. If $`q𝒰H`$ then one finds $`\alpha ,\beta 𝒩H`$ with $`q=\alpha \beta ^1`$ by \[6, Theorem 1 and proof of Theorem 10\] or \[24, 2.7\]. Applying the same to $`q^{}`$ we may as well achieve $`q=\beta ^1\alpha `$. Apply this now inductively to $`q_1=\beta _1^1\alpha _1`$, $`\beta _1q_2=\beta _2^1\alpha _2,\mathrm{}`$ to get the non-zero divisor $`\beta =\beta _1\mathrm{}\beta _k𝒩H`$ (which is invertible in $`𝒰H`$) such that $`\beta q_i𝒩H`$ for $`i=1,\mathrm{},k`$. It follows from \[16, Theorem 4\] that $`\beta x`$ is not a zero divisor in $`𝒩G`$, therefore it becomes invertible in $`𝒰G`$. ∎ ###### 3.6 Lemma. Let $`DG`$ be the division closure of $`KG`$ in $`𝒰G`$, where $`K=\overline{K}`$. If $`\alpha \mathrm{Aut}(G)`$ then $$\alpha (DG)=DG^{}=DG.$$ ###### Proof. Of course $`\alpha (DG)`$ is the division closure of $`\alpha (KG)`$ in $`\alpha (𝒰G)`$. Since $`\alpha (KG)=KG`$ and $`\alpha (𝒰G)=𝒰G`$ we conclude $`\alpha (DG)=DG`$. The proof for the anti-automorphism $``$ is identical. Here $`KG=KG^{}`$ since $`K`$ is closed under complex conjugation. ∎ ###### 3.7 Definition. Mainly for proofs by induction we will use the following constructions of classes of groups in addition to Definition 1.10 and Definition 1.4: * For a class $`𝒳`$ of groups, $`L𝒳`$ denotes the class of all groups which are locally of class $`𝒳`$, i.e. every finitely generated subgroup belongs to $`𝒳`$. Such groups are directed unions of groups in $`𝒳`$. * For two classes of groups $`𝒳`$ and $`𝒴`$, the class $`𝒳𝒴`$ consists of all groups $`G`$ with normal subgroup $`H𝒳`$ and quotient $`G/H𝒴`$. Let $`𝒳^\times `$ consist of all finite direct sums of groups in $`𝒳`$. * $`()`$ is the class of free groups, $`(𝒩)`$ is the class of finite groups, and $``$ is the class of finitely generated virtually abelian groups. * The elementary amenable groups are denoted by $`𝒜`$. This is the smallest class of groups which contains all abelian and all finite groups and is closed under extensions and directed unions. ###### Proof of Proposition 3.1. Case 1: $`G/H`$ finitely generated free abelian. By induction we can immediately reduced to the case where $`G/H`$ is infinite cyclic. Since $`DH`$ is a skew field an arbitrary non-zero element of $`DHG/H`$ has after multiplication with units in $`DHG/H`$ the shape $`1+q_1t+q_2t^2+\mathrm{}q_kt^k`$ where $`t`$ is a generator of $`G/H`$. Lemma 3.5 implies therefore that each element of $`(DH)G\{0\}`$ becomes invertible in $`𝒰G`$, in particular $`(DH)G`$ has no zero divisors. Therefore its Ore completion is a skew field which embeds into $`𝒰G`$. It embeds into $`DG`$ and is division closed, therefore both coincide, $`DG`$ is a skew field and by Lemma 3.4 the strong Atiyah conjecture holds for $`G`$. If, as a next step, $`G/H`$ is finitely generated free abelian, induction immediately implies that the strong Atiyah conjecture holds for $`G`$. Case 2: $`G/H`$ finitely generated virtually abelian. It has a finitely generated free abelian normal subgroup $`A_0`$. Its inverse image $`G_0`$ in $`G`$ fulfills the strong Atiyah conjecture by the previous step. Since $`H`$ fulfills the strong Atiyah conjecture, by Lemma 3.4 the division closure $`DH`$ of $`KH`$ in $`𝒰H`$ is a skew field. Conjugation with an element of $`G`$ gives an automorphism of $`H`$. Using Lemma 3.6, by \[17, 2.1\] the subring $`(DH)G`$ generated by $`DH`$ and $`G`$ is a crossed product: $`(DH)G=DH(G/H)`$. The same applies to every finite extension $`HH_E`$, therefore $`(DH)H_E=DH(H_E/H)`$. By \[17, 4.2\] for these finite extensions $`(DH)H_E=DH_E`$, since $`DH`$ is a skew field and hence Artinian. By assumption all the groups $`H_E`$ are torsion-free and fulfill the strong Atiyah conjecture. By Lemma 3.4 $`DH_E=DHE`$ then is a skew field for every finite $`E<G/H`$. Because of \[18, 4.5\], $`DHG/H=(DH)G`$ is an Ore domain. The proof shows that its Ore completion (a skew field!) is obtained by inverting $`DH(G_0/H)\{0\}`$. Since $`DH(G_0/H)DG_0`$, and the latter is a skew field contained in $`𝒰G`$, all inverses $`\left(DH(G_0/H)\{0\}\right)^1`$ are contained in $`𝒰G`$. Therefore the same is true for the Ore completion of $`DHG/H`$, which is a skew field and as a result coincides with the division closure $`DG`$. By Lemma 3.4 the strong Atiyah conjecture holds for $`G`$. Case 3: $`A`$ elementary amenable. This is done by transfinite induction, where we use roughly the description of elementary amenable groups given in \[13, Section 3\]: Let $`𝒜_0`$ be the class of finite groups, and for an ordinal $`\alpha `$ set $`𝒜_{\alpha +1}:=(L𝒜_\alpha )`$. If $`\alpha `$ is a limit ordinal set $`𝒜_\alpha =_{\beta <\alpha }𝒜_\beta `$. Then $`𝒜=_{\alpha 0}𝒜_\alpha `$. One can easily show, as in \[17, 4.9\], that each $`𝒜_\alpha `$ and $`L𝒜_\alpha `$ is subgroup closed and closed under extensions with finite quotient. The case $`A𝒜_0`$ is trivial, because then the conclusion is part of the assumptions. Let $`\alpha `$ be the least ordinal with $`A𝒜_\alpha `$. Then $`\alpha =\beta +1`$. Assume by induction that the statement is true for $`𝒜_\beta `$. If $`A=_{iI}A_i`$ with $`A_i𝒜_\beta `$, then $`G=G_i`$ where $`G_i`$ is the inverse image of $`A_i`$ in $`G`$, i.e. an extension of $`H`$ by $`G_i`$. By induction and Lemma 2.3 the Atiyah conjecture is true for $`G`$, therefore it is true if the quotients belong to $`L𝒜_\beta `$. Let now $`A`$ be an extension with kernel $`BL𝒜_\beta `$ and quotient $`Q`$ which is finitely generated virtually abelian. We want to apply the induction hypothesis and Case 2 to $`1G_1GQ1`$, where for a finite subgroup $`E`$ of $`Q`$ we let $`G_E`$ be the inverse image of $`E`$ in $`G`$ (in particular $`G_1`$ is the inverse image of $`B`$). To apply the results of Case 2 we have to establish the Atiyah conjecture for all the groups $`G_E`$. We obtain exact sequences $`1HG_EB_E1`$ where $`B_E`$ is a finite extension of $`B`$. Hence $`B_EL𝒜_\beta `$. Note that $`B_E`$ is a subgroup of $`A`$, therefore each finite subgroup of $`B_E`$ is a finite subgroup of $`A`$, too. This implies that the assumptions of the proposition are fulfilled for $`1HG_EB_E1`$ and therefore $`G_E`$ fulfills the strong Atiyah conjecture by induction. Hence the assumptions of Case 2 are fulfilled for $`1G_1GQ1`$ and the result follows. ∎ In the course of the proof we also established: ###### 3.8 Lemma. If, in the situation of Proposition 3.1, $`G/H`$ is finitely generated virtually abelian then the ring $`DG`$ is an Ore localization of $`DHG/H`$. ###### 3.9 Example. Let $`G𝒢`$ (as defined in Definition 1.4) be a torsion-free group which fulfills the strong Atiyah conjecture with $`K=\overline{K}`$, and $`f:GG`$ a group homomorphism. Then the mapping torus group $$T_f:=<t,G|tgt^1=f(g),gh=(gh)g,hG>$$ fulfills the strong Atiyah conjecture over $`T_f`$. If $`f`$ is injective, then $`T_f`$ fulfills the strong Atiyah conjecture even over $`KT_f`$ and we don’t need the assumption $`G𝒢`$. ###### Proof. Let $`E_f`$ be the direct limit of $`G\stackrel{f}{}G\stackrel{f}{}G\mathrm{}`$. Then $`E_f`$ fulfills the strong Atiyah conjecture by Proposition 2.4 or, if $`f`$ is injective, by Proposition 2.3. The claim follows from Proposition 3.1 because $`T_f`$ is an extension of $`E_f`$ with quotient $``$. ∎ ## 4 Proofs of the Atiyah conjecture In this section we prove Theorem 1.11 and Proposition 1.13 and give additional information about $`𝒟`$. ###### 4.1 Lemma. Let $`H=F_1\times F_2\times \mathrm{}`$ be a direct sum of non-abelian free groups. Let $`HG`$ be a torsion-free finite extension of $`H`$. Then every finitely generated subgroup of $`G`$ is contained in a group $`V`$ which fits into an extension $`1UVA1`$ where $`U`$ is a subgroup of a direct sum of free groups and $`A`$ is torsion-free elementary amenable. ###### Proof. By \[18, 13.2\] every finitely generated subgroup of $`G`$ is contained in a finite extension of a finitely generated product of free groups. So we may assume that $`G`$ (and $`H`$) are finitely generated. We now use the ideas of \[18, 13.4\] where Linnell proves a related result. For a group $`V`$ let $`S_nV`$ be the intersection of normal subgroups of index $`n`$. Note that this is a characteristic subgroup, and if $`V`$ is finitely generated, $`S_nV`$ has finite index in $`S`$ by \[25, 8.2.7\]. Set $`H_n:=S_nF_1\times S_nF_2\mathrm{}S_nF_k`$. Let $`H_n^{}`$ be the commutator subgroup of $`H_n`$. Then $`H/H_n^{}`$ is virtually abelian and torsion-free: it is the direct sum of $`F_i/(S_nF_i)^{}`$ containing the abelian subgroup $`S_nF_i/(S_nF_i)^{}`$ of finite index, and whenever $`F`$ is free and $`NF`$ is a normal subgroup then $`F/N^{}`$ is torsion free (compare e.g. the last paragraph in the proof of 3.4.8 of ). Since $`G`$ is torsion-free, by \[18, 13.4\] for every prime number $`p`$ there is a subgroup $`K_p<H`$ which is normal in $`G`$ such that $`G/K_p`$ is virtually abelian and does not contain a finite subgroup of index $`p`$. For $`n_p`$ sufficiently large $`H_{n_p}^{}<K_p`$. We do this for all prime factors of $`\left|G/H\right|`$ and let $`n`$ be the maximum of the $`n_p`$ (which exists since $`G/H`$ is finite). Then $`1H_n^{}GA1`$ is an extension where $`A`$ is virtually abelian. We also have an extension $`1H/H_n^{}AG/H1`$. Since $`H/H_n^{}`$ is torsion-free, every torsion element of $`A`$ is mapped to an element of the same order in $`G/H`$. Therefore it suffices to check for a prime factor $`p`$ of $`\left|G/H\right|`$ and $`xA`$ with $`x^p=1`$ that $`x=1`$. But we also have the extension $`1K_p/H_n^{}AG/K_p1`$ where $`K_p/H_n^{}<H/H_n^{}`$ is torsion-free and $`G/K_p`$ does not contain $`p`$-torsion. This concludes the proof. ∎ ###### Proof of Theorem 1.11 and Corollary 1.12. Remember that every group in $`𝒟`$ is torsion-free and belongs by Lemma 2.6 to $`𝒢`$. The result follows immediately from Proposition 2.2, Proposition 2.4, and Proposition 3.1. The corollary is an immediate consequence of the theorem (because of Lemma 3.4). ∎ ###### Proof of Proposition 1.13. By , every finitely generated free group is residually torsion-free nilpotent. Consequently, the same is true for every direct sum of free groups. Therefore, these groups are subgroups of inverse limits of torsion-free nilpotent (and therefore elementary amenable) groups. It follows that direct sums of arbitrary free groups (as directed unions of sums of finitely generated ones) and subgroups of these belong to $`𝒟`$. By Lemma 4.1, every torsion-free finite extension of a direct sum of free groups then also belongs to $`𝒟`$. It now follows immediately from the definition of $`𝒞^{}`$ that every torsion-free element of $`𝒞^{}`$ belongs to $`𝒟`$. Consider (for every ordinal $`\alpha `$) the subclasses $`𝒞_\alpha `$ of $`𝒞`$ with $$𝒞_0:=()(𝒩)\text{and}𝒞_{\alpha +1}:=(L𝒞).$$ If $`\alpha `$ is a limit ordinal, set $`𝒞_\alpha :=_{\beta <\alpha }𝒞_\beta `$. By induction, we proof that every torsion-free element of $`𝒞_\alpha `$ belongs to $`𝒟`$. We have just shown this for $`\alpha =0`$. If $`\alpha `$ is a limit ordinal, nothing is to proof. By \[17, Lemma 4.9\], each $`L𝒞_\alpha `$ is closed under extensions with finite quotients. If $`G𝒞_{\alpha +1}`$ is torsion-free, we have an extension $`1HG\stackrel{𝑝}{}A1`$ with $`A`$ elementary amenable (since $`A`$ is abelian by finite) and $`HL𝒞_\alpha `$. If $`E<A`$ is finite, therefore $`p^1(E)L𝒞_\alpha `$ (as a finite extension of $`H`$). Moreover, $`p^1(E)`$ is torsion-free, since it is a subgroup of the torsion-free group $`G`$. By induction, $`p^1(E)𝒟`$. It follows that $`G𝒟`$. This finishes the induction step. Since $`𝒞=_\alpha 𝒞_\alpha `$ (by \[17, Lemma 4.9\]), the statement about $`𝒞`$ follows. Next, we prove that $`𝒟`$ is closed under direct sums. Since every direct sum is a union of finite direct sums, by induction it suffices to consider direct sums with two summands. Fix $`G𝒟`$. The start of the induction is the trivial observation that $`\{1\}\times G𝒟`$. For the induction step, we have to check that $`H\times G𝒟`$ in the following three cases: 1. There is an exact sequence $`1UH\stackrel{𝑝}{}A1`$ with $`A`$ elementary amenable and such that $`p^1(E)\times G𝒟`$ for every finite subgroup $`E`$ of $`A`$. But then we have an exact sequence $$1U\times GH\times G\stackrel{𝑞}{}A1$$ and $`q^1(E)=p^1(E)\times G`$ for every finite subgroup $`E`$ of $`A`$, and it follows that $`H\times G𝒟`$. 2. $`H`$ is a direct or inverse limit of a directed system of groups $`H_i`$, and $`H_i\times G𝒟`$ for every $`i`$. Then $`H\times G`$ is the direct or inverse limit of the directed system $`H_i\times G`$ by Lemma 4.2, and therefore $`H\times G𝒟`$. 3. If $`HU`$ and $`U\times G𝒟`$, then $`H\times GU\times G`$ which implies $`H\times G𝒟`$. The detailed construction of families $`𝒟_\alpha `$ such that this proof applies is standard (compare the similar construction of $`𝒞_\alpha `$ used above) and is left to the reader. Every direct product is the inverse limit of finite direct sums. It follows that $`𝒟`$ is also closed under taking direct products. We now proof by induction that $`𝒟`$ is closed under free products with (arbitrary) free groups. The induction starts with the observation that $`\{1\}F𝒟`$ if $`F`$ is free, what we have already checked. For the induction step, we again have to check three cases. 1. Assume $`HU`$ and $`UF𝒟`$ for every free group $`F`$. Then $`HFUF`$ and consequently $`HF𝒟`$. 2. Assume we have an exact sequence $`1UHA\stackrel{𝑝}{}1`$ with $`A`$ elementary amenable and $`p^1(E)F𝒟`$ for every free group $`F`$ and every finite subgroup $`E`$ of $`A`$. Consider the projection $`q:=p1:HFA`$. Then for every $`E<A`$ Lemma 4.3 implies $`q^1(E)p^1(E)F^{}`$ with some possibly different free group $`F^{}`$. Since the induction hypothesis applies to $`p^1(E)F^{}`$ if $`E`$ is finite, it follows that $`HF𝒟`$. 3. Assume $`H`$ is the direct or inverse limit of a directed system of groups $`H_i`$, and $`H_iF𝒟`$ for every free group $`F`$. By \[26, Lemma 2.9 and 2.10\] $`HF`$ is contained in the direct or inverse limit of the directed system $`H_iF`$. It follows that $`HF𝒟`$. For arbitrary free products observe that $`GH`$ is a subgroup of $`(G\times H)(G\times H)`$ which is a subgroup of $`(G\times H)`$ (it is contained in the kernel of the projection onto $``$ by \[11, Lemma 5.5\]). Since $`𝒟`$ is subgroup closed, $`GH𝒟`$ if $`G,H𝒟`$ follows from what we have proved about direct sums and free products with free groups. The class $`𝒟`$ is closed under arbitrary free products by induction and considering directed unions (an arbitrary free product is the directed union of finite free products). This finishes the proof. ∎ The following lemmas were needed in the proof of Proposition 1.13. ###### 4.2 Lemma. If $`\pi `$ is the direct or inverse limit of a system of groups $`\pi _i`$ and if $`G`$ is any group, then $`\pi \times G`$ is the direct or inverse limit of $`\pi _i\times G`$. ###### Proof. Compare e.g. in \[26, 2.8 and 2.9\]. ∎ ###### 4.3 Lemma. Assume $`p:HA`$ a homomorphism of groups. Consider $`q:=p1:HGA`$. Then, for every subgroup $`E`$ of $`A`$, $`q^1(E)=p^1(E)(_{tT}(t^1Gt))`$, where $`T`$ is a transversal for $`p^1(E)`$ in $`H`$, i.e. a system of representatives for the cosets $`H/p^1(E)`$. ###### Proof. Obviously, $`q`$ maps $`p^1(E)`$ and $`t^1Gt`$ to $`E`$. On the other hand, a normal form argument implies that each element in $`q^1(E)`$ lies in the subgroup of $`HG`$ generated by $`p^1(E)`$ and $`t^1Gt`$ ($`tT`$). The normal form argument also implies that this subgroup is isomorphic to the free product $`p^1(E)(_{tT}(t^1Gt))`$. ∎ We now explain how to fill the gap (pointed out in Section 1) of Linnell’s result about the Atiyah conjecture for the class $`𝒞^{}`$, using our methods (and obtaining a slightly weaker conclusion). ###### 4.4 Corollary. Assume $`G𝒞^{}`$ and there is a bound on the orders of the finite subgroups of $`G`$. Then the strong Atiyah conjecture is true over $`G`$. ###### Proof. Linnell’s proof in \[18, Section 13\] applies as soon as the strong Atiyah conjecture over $`G`$ is established for every direct sum of free groups $`G`$. However, we have seen that these groups belong to $`𝒟`$ and therefore satisfy the strong Atiyah conjecture over $`G`$. ∎ ## 5 Generalizations In the second part of the paper we concentrated on torsion-free groups. By Linnell’s methods, many of the results hold if instead of this one requires a bound on the orders of finite subgroups. In particular, a version of Proposition 3.1 remain true if $`Q`$ is not assumed to be torsion-free but only has a bound on the orders of its finite subgroups. One proves this by transfinite induction, using a little bit more of the machinery of Linnell . Correspondingly, one can then (similarly to $`𝒟`$) define a larger class of groups (not necessarily torsion-free) for which the strong Atiyah conjecture is true. Another possible generalization is given by only considering the $`1`$-dimensional Atiyah conjecture. This is defined as in Definition 1.2, but one does consider only $`1\times 1`$-matrices $`A`$, i.e. $`AKG`$. This is interesting because already the strong $`1`$-dimensional Atiyah conjecture for torsion-free groups implies the zero divisor conjecture. Moreover, Linnell proves that the $`1`$-dimensional strong Atiyah conjecture for torsion-free groups is stable under extensions with right-orderable groups. One immediately checks that there is a $`1`$-dimensional versions of Proposition 2.4. Hence we can define a class of groups similar to $`𝒟`$ but which is closed under extensions with right-orderable quotient (instead of extensions with elementary amenable quotient), and the $`1`$-dimensional strong Atiyah conjecture over $`G`$ is true if $`G`$ is contained in this class. In particular, $`G`$ is zero divisor free. ## 6 Treelike groups ###### 6.1 Proposition. Let $`H`$ be a finite group. It follows that $`H`$ fulfills the Atiyah conjecture of order $`\frac{1}{L}`$ over $`KH`$, with $`L=\left|H\right|`$. Let $`G`$ be a group and $`\mathrm{\Omega }`$ and $`\mathrm{\Delta }`$ sets with commuting $`G`$-action from the left and $`H`$-action from the right such that $`\mathrm{\Omega }`$ and $`\mathrm{\Delta }`$ are free $`H`$-sets. Let $`\mathrm{\Omega }=\mathrm{\Omega }^{}X`$ and assume that $`\mathrm{\Delta }`$ and $`\mathrm{\Omega }^{}`$ are free $`G`$-sets and $`X`$ consists of $`1r<\mathrm{}`$ $`H`$-orbits. Assume that there is a bijective map $`\varphi :\mathrm{\Delta }\mathrm{\Omega }`$ which is an $`H`$-almost $`G`$-map and a right $`H`$-blockwise map (compare Definition 6.2). Then $`G`$ fulfills the Atiyah conjecture of order $`\frac{1}{rL}`$ over $`KG`$. We used the following notation: ###### 6.2 Definition. The map $`\varphi :\mathrm{\Delta }\mathrm{\Omega }`$ is an *$`H`$-almost $`G`$-map* if for each $`gG`$ the set $`\{x\mathrm{\Delta }\varphi (gx)g\varphi (x)\}`$ is contained in only finitely many $`H`$-orbits. It is called a *right $`H`$-blockwise* map if each $`H`$-orbit of $`\mathrm{\Delta }`$ is mapped bijectively to an $`H`$-orbit of $`\mathrm{\Omega }`$. ###### 6.3 Remark. Whenever the assumptions of Proposition 6.1 are fulfilled, we can replace $`H`$ by $`\{1\}`$ (and therefore get $`L=1`$). However, $`X`$ then consists of $`r\times \left|H\right|=rL`$ $`\{1\}`$-orbits (i.e. points), and the conclusion is unchanged. ###### Proof of Proposition 6.1. Our proof generalizes Linnell’s approach in . Assume that we have the $`G`$-$`H`$ sets $`\mathrm{\Omega }`$ and $`\mathrm{\Delta }`$ as above. Given $`AM_n(KG)`$, the $`G`$-operation induces bounded operators $`A_\mathrm{\Delta }`$ on $`l^2(\mathrm{\Delta })^n`$ and $`A_\mathrm{\Omega }^{}`$ on $`l^2(\mathrm{\Omega }^{})^n`$ in the obvious way. Set $`A_\mathrm{\Omega }:=A_\mathrm{\Omega }^{}0`$ on $`l^2(\mathrm{\Omega })^n=l^2(\mathrm{\Omega }^{})^nl^2(X)^n`$. Let $`PM_n(𝒩G)`$ be the projection onto the image of $`A`$. Since $`\mathrm{\Delta }`$ and $`\mathrm{\Omega }^{}`$ are free $`G`$-sets one can extend the above construction to $`P`$. It follows immediately that $`P_\mathrm{\Delta }`$ is the projection onto the image of $`A_\mathrm{\Delta }`$ and $`P_\mathrm{\Omega }`$ the projection onto the image of $`A_\mathrm{\Omega }`$. Since the $`H`$-actions commute with the $`G`$-actions, all operators constructed in this way are $`H`$-equivariant. There are only finitely many $`gG`$ such that $`A=:(a_{ij})M_n(G)`$ has a non-trivial coefficient of $`g`$ in at least one of the $`a_{ij}`$. For each of these, there are only finitely many $`H`$-orbits in $`\mathrm{\Delta }`$ on which $`\varphi (gx)g\varphi (x)`$. Let $`\mathrm{\Delta }_0`$ be the union of this finite number of $`H`$-orbits in $`\mathrm{\Delta }`$, and set $`\mathrm{\Delta }_c=\mathrm{\Delta }\mathrm{\Delta }_0`$ and $`\mathrm{\Omega }_c:=\varphi (\mathrm{\Delta }_c)`$. Restricted to $`l^2(\mathrm{\Delta }_c)^n`$, $`\varphi ^{}A_\mathrm{\Omega }\varphi `$ and $`A_\mathrm{\Delta }`$ coincide: $$A_\mathrm{\Delta }|_{l^2(\mathrm{\Delta }_c)^n}=\varphi ^{}A_\mathrm{\Omega }\varphi |_{l^2(\mathrm{\Delta }_c)^n}.$$ (6.4) Choose an exhaustion $`\mathrm{\Delta }_0\mathrm{\Delta }_1\mathrm{\Delta }_2\mathrm{}\mathrm{\Delta }`$ of $`\mathrm{\Delta }`$ (i.e. $`\mathrm{\Delta }=_k\mathrm{\Delta }_k`$) where each $`\mathrm{\Delta }_k`$ consists of finitely many (entire) $`H`$-orbits. Then (since $`\varphi `$ is an $`H`$-blockwise map) $`\mathrm{\Omega }_k:=\varphi (\mathrm{\Delta }_k)`$ also consists of finitely many $`H`$-orbits. If $`\mathrm{\Delta }`$ consists of uncountably many $`H`$-orbits, we have to use an exhaustion index by an uncountable directed index set. The proof remains essentially unchanged. Since our applications only deal with the countable setting for simplicity we stick to the index set $``$. As $`\mathrm{\Delta }=_k\mathrm{\Delta }_k`$ we have $`A_\mathrm{\Delta }(l^2(\mathrm{\Delta })^n)=\overline{_kA_\mathrm{\Delta }(l^2(\mathrm{\Delta }_k)^n)}`$. Let $`P_{\mathrm{\Delta },k}`$ be the projection onto $`\overline{A_\mathrm{\Delta }(l^2(\mathrm{\Delta }_k)^n)}`$, and $`P_{\mathrm{\Omega },k}`$ the projection onto $`\overline{A_\mathrm{\Omega }(l^2(\mathrm{\Omega }_k)^n)}`$. Because in the matrix $`A=(a_{ij})`$ only finitely many elements of $`G`$ have non-trivial coefficient in at least one of the $`a_{ij}`$, $`A`$ has finite propagation, i.e. for every $`k`$ we find $`n(k)`$ such that $`A_\mathrm{\Delta }(l^2(\mathrm{\Delta }_k)^n)l^2(\mathrm{\Delta }_{n(k)})^n`$. Similarly (if $`n(k)`$ is chosen big enough) $`A_\mathrm{\Omega }(l^2(\mathrm{\Omega }_k)^n)l^2(\mathrm{\Omega }_{n(k)})^n`$. Particularly, $`A_\mathrm{\Delta }(l^2(\mathrm{\Delta }_k)^n)`$ and $`A_\mathrm{\Omega }(l^2(\mathrm{\Omega }_k)^n)`$ are the ranges of finite matrices over $`KH`$ to which the Atiyah conjecture applies and by assumption on $`H`$ $$\begin{array}{cc}\hfill \mathrm{tr}_H(P_{\mathrm{\Delta },k})& =dim_H(A_\mathrm{\Delta }(l^2\mathrm{\Delta }_k)^n)\frac{1}{L},\hfill \\ \hfill \mathrm{tr}_H(P_{\mathrm{\Omega },k})& =dim_H(A_\mathrm{\Omega }(l^2\mathrm{\Omega }_k)^n)\frac{1}{L}.\hfill \end{array}$$ (6.5) Let $`P_{\mathrm{\Delta },k,c}`$ be the projection onto $`A_\mathrm{\Delta }(l^2(\mathrm{\Delta }_k\mathrm{\Delta }_c)^n)`$ and $`P_{\mathrm{\Delta },c}`$ the projection onto $`A_\mathrm{\Delta }(l^2(\mathrm{\Delta }_c)^n)`$. Define $`P_{\mathrm{\Omega },k,c}`$ and $`P_{\mathrm{\Omega },c}`$ correspondingly. By Equation (6.4) $`P_{\mathrm{\Delta },k,c}=\varphi ^{}P_{\mathrm{\Omega },k,c}\varphi `$ and $`P_{\mathrm{\Delta },c}=\varphi ^{}P_{\mathrm{\Omega },c}\varphi `$. In particular, if $`x\mathrm{\Delta }^nl^2(\mathrm{\Delta })^n`$ we have (with $`\varphi `$ diagonally extended to $`\mathrm{\Delta }^n`$) $$\begin{array}{cc}\hfill P_{\mathrm{\Delta },k,c}x,x_{l^2(\mathrm{\Delta })^n}& =P_{\mathrm{\Omega },k,c}\varphi (x),\varphi (x)_{l^2(\mathrm{\Omega })^n},\hfill \\ \hfill P_{\mathrm{\Delta },c}x,x_{l^2(\mathrm{\Delta })^n}& =P_{\mathrm{\Omega },c}\varphi (x),\varphi (x)_{l^2(\mathrm{\Omega })^n}.\hfill \end{array}$$ (6.6) Set $`Q_{\mathrm{\Delta },k}:=P_{\mathrm{\Delta },k}P_{\mathrm{\Delta },k,c}`$, $`Q_\mathrm{\Delta }:=P_\mathrm{\Delta }P_{\mathrm{\Delta },c}`$, and define $`Q_{\mathrm{\Omega },k}`$ and $`Q_\mathrm{\Omega }`$ correspondingly. These are the projections onto the complement of the smaller sets in the larger ones. Note that the images of the $`Q`$’s are already obtained from $`l^2(\mathrm{\Delta }_0)^n`$ and $`l^2(\mathrm{\Omega }_0)^n`$, respectively. More precisely, let $`A_0`$ be the restriction of $`A_\mathrm{\Delta }`$ to $`l^2(\mathrm{\Delta }_0)^n`$. Then $`Q_{\mathrm{\Delta },k}A_0`$ induces a weakly exact sequence (i.e. the images are dense in the kernels) $$0\mathrm{ker}(Q_{\mathrm{\Delta },k}A_0)l^2(\mathrm{\Delta }_0)^n\stackrel{Q_{\mathrm{\Delta },k}A_0}{}\mathrm{im}(Q_{\mathrm{\Delta },k})0.$$ (6.7) We only have to check weak exactness at $`\mathrm{im}(Q_{\mathrm{\Delta },k})`$. $`Q_{\mathrm{\Delta },k}(A(l^2(\mathrm{\Delta })^k))`$ is dense in $`\mathrm{im}(Q_{\mathrm{\Delta },k})`$. But $`A(l^2(\mathrm{\Delta }_k)^n)=A(l^2(\mathrm{\Delta }_0)^n)+A(l^2(\mathrm{\Delta }_k\mathrm{\Delta }_c)^n)`$ and $`Q_{\mathrm{\Delta },k}(A(l^2(\mathrm{\Delta }_k\mathrm{\Delta }_c)^n))=0`$. For the kernel we have the weakly exact sequence $$\begin{array}{c}0\mathrm{ker}(A_0)\mathrm{ker}(Q_{\mathrm{\Delta },k}A_0)\stackrel{A_0}{}\hfill \\ \hfill \overline{A_\mathrm{\Delta }(l^2(\mathrm{\Delta }_0)^n)A_\mathrm{\Delta }(l^2(\mathrm{\Delta }_k\mathrm{\Delta }_c)^n)}0,\end{array}$$ (6.8) which would be clear with $`A_\mathrm{\Delta }(l^2(\mathrm{\Delta }_k\mathrm{\Delta }_c)^n)`$ replaced by $`\mathrm{ker}(Q_{\mathrm{\Delta },k})`$ —but if $`fA_\mathrm{\Delta }(l^2(\mathrm{\Delta }_k)^n)`$ and in particular if $`fA_\mathrm{\Delta }(l^2(\mathrm{\Delta }_0)^n)`$, then $`f\mathrm{ker}(Q_{\mathrm{\Delta },k})`$ if and only if $`fA_\mathrm{\Delta }(l^2(\mathrm{\Delta }_k\mathrm{\Delta }_c)^n)`$. This is true since all the vector spaces we are considering here are finite dimensional (a consequence of the assumption that $`H`$ is finite). In the same way one gets the weakly exact sequences $$0\mathrm{ker}(Q_\mathrm{\Delta }A_0)l^2(\mathrm{\Delta }_0)\stackrel{Q_\mathrm{\Delta }A_0}{}\mathrm{im}(Q_\mathrm{\Delta })0,$$ (6.9) $$0\mathrm{ker}(A_0)\mathrm{ker}(Q_\mathrm{\Delta }A_0)\stackrel{A_0}{}\overline{A_\mathrm{\Delta }(l^2(\mathrm{\Delta }_0)^n)A_\mathrm{\Delta }(l^2(\mathrm{\Delta }_c)^n)}0.$$ (6.10) Observe that $$\overline{A_\mathrm{\Delta }(l^2(\mathrm{\Delta }_0)^n)A_\mathrm{\Delta }(l^2(\mathrm{\Delta }_c)^n)}=\overline{\underset{k}{}A_\mathrm{\Delta }(l^2(\mathrm{\Delta }_0)^n)A_\mathrm{\Delta }(l^2(\mathrm{\Delta }_k\mathrm{\Delta }_c)^n)},$$ hence because of continuity of $`dim_H`$ \[20, 1.4\] $$\begin{array}{c}dim_H(\overline{A_\mathrm{\Delta }(l^2(\mathrm{\Delta }_0)^n)A_\mathrm{\Delta }(l^2(\mathrm{\Delta }_c)^n)})\hfill \\ \hfill =\underset{k\mathrm{}}{lim}dim_H(\overline{A_\mathrm{\Delta }(l^2(\mathrm{\Delta }_0)^n)A_\mathrm{\Delta }(l^2(\mathrm{\Delta }_k\mathrm{\Delta }_c)^n)}).\end{array}$$ (6.11) Using the additivity of $`dim_H`$ under short weakly exact sequences \[20, 1.4\], Equation (6.11) implies together with Equation (6.7), (6.8), (6.9), and (6.10) (where all numbers are finite and bounded by $`dim_H(l^2(\mathrm{\Delta }_0)^n)`$) $$\begin{array}{c}\mathrm{tr}_H(Q_\mathrm{\Delta })=dim_H(\mathrm{im}(Q_\mathrm{\Delta }))=\underset{k\mathrm{}}{lim}dim_H(\mathrm{im}(Q_{\mathrm{\Delta },k}))\hfill \\ \hfill =\underset{k\mathrm{}}{lim}\mathrm{tr}_H(Q_{\mathrm{\Delta },k}).\end{array}$$ (6.12) Exactly in the same way one obtains $$\mathrm{tr}_H(Q_\mathrm{\Omega })=\underset{k\mathrm{}}{lim}\mathrm{tr}_H(Q_{\mathrm{\Omega },k}).$$ (6.13) Choose a set of representatives $`T`$ for the $`H`$-orbits of $`\mathrm{\Delta }`$. Because $`\varphi `$ is an $`H`$-blockwise bijective map, $`\varphi (T)`$ will be a set of representatives for the $`H`$-orbits of $`\mathrm{\Omega }`$. For $`tT`$ set $`t_i:=(0,\mathrm{},0,t,0,\mathrm{},0)\mathrm{\Delta }^n`$ ($`t`$ at the $`i`$-th position). Then (by the definition of the $`H`$-trace) $$\begin{array}{cc}\hfill \mathrm{tr}_H(Q_\mathrm{\Delta })=& \underset{tT}{}\underset{i=1}{\overset{n}{}}Q_\mathrm{\Delta }t_i,t_i_{l^2(\mathrm{\Delta })},\hfill \\ \hfill \mathrm{tr}_H(Q_\mathrm{\Omega })=& \underset{tT}{}\underset{i=1}{\overset{n}{}}Q_\mathrm{\Omega }\varphi (t_i),\varphi (t_i)_{l^2(\mathrm{\Delta })}.\hfill \end{array}$$ Because $`\mathrm{\Delta }`$ is a free $`G`$-set, $`Gt\mathrm{\Delta }`$ can for each $`tT`$ be identified with $`G`$ and therefore $$\underset{i=1}{\overset{n}{}}P_\mathrm{\Delta }t_i,t_i_{l^2(\mathrm{\Delta })^n}=\mathrm{tr}_G(P)=dim_G(\mathrm{im}(A)).$$ Similarly, since $`\mathrm{\Omega }^{}`$ is a free $`G`$-set, if $`\varphi (t)\mathrm{\Omega }^{}`$ then $$\begin{array}{c}\underset{i=1}{\overset{n}{}}P_\mathrm{\Omega }\varphi (t_i),\varphi (t_i)_{l^2(\mathrm{\Omega })^n}=\underset{i=1}{\overset{n}{}}P_\mathrm{\Omega }\varphi (t_i),\varphi (t_i)_{l^2(\mathrm{\Omega }^{})^n}=\hfill \\ \hfill \mathrm{tr}_G(P)=dim_G(\mathrm{im}(A)).\end{array}$$ However, if $`\varphi (t)X`$ then by the construction of $`P_\mathrm{\Omega }`$ $$P_\mathrm{\Omega }\varphi (t_i),\varphi (t_i)_{l^2(\mathrm{\Omega })}=0.$$ Therefore $$\underset{tT}{}\underset{i=1}{\overset{n}{}}(P_\mathrm{\Delta }t_i,t_i_{l^2(\mathrm{\Delta })^n}P_\mathrm{\Omega }\varphi (t_i),\varphi (t_i)_{l^2(\mathrm{\Omega })^n})=rdim_G(\mathrm{im}(A))$$ (6.14) (where $`r=\left|\varphi (T)X\right|`$ is the number of $`H`$-orbits in $`X`$). Note that the sum is finite so that convergence is not an issue here. We want to use this equation to determine the value of $`dim_G(\mathrm{im}(A))`$. From our splitting of $`P_\mathrm{\Delta }`$ and $`P_\mathrm{\Omega }`$ we see that for each $`tT`$ and $`i\{1,\mathrm{},n\}`$ $$\begin{array}{cc}& P_\mathrm{\Delta }t_i,t_i_{l^2(\mathrm{\Delta })^n}P_\mathrm{\Omega }\varphi (t_i),\varphi (t_i)_{l^2(\mathrm{\Omega })^n}\hfill \\ & =\underset{\stackrel{(\text{6.6})}{=}0}{\underset{}{P_{\mathrm{\Delta },c}t_i,t_iP_{\mathrm{\Omega },c}\varphi (t_i),\varphi (t_i)}}+Q_\mathrm{\Delta }t_i,t_iQ_\mathrm{\Omega }\varphi (t_i),\varphi (t_i).\hfill \end{array}$$ (6.15) In the same way $$\begin{array}{cc}& P_{\mathrm{\Delta },k}t_i,t_i_{l^2(\mathrm{\Delta })^n}P_{\mathrm{\Omega },k}\varphi (t_i),\varphi (t_i)_{l^2(\mathrm{\Omega })^n}\hfill \\ & =\underset{\stackrel{(\text{6.6})}{=}0}{\underset{}{P_{\mathrm{\Delta },k,c}t_i,t_iP_{\mathrm{\Omega },k,c}\varphi (t_i),\varphi (t_i)}}+Q_{\mathrm{\Delta },k}t_i,t_iQ_{\mathrm{\Omega },k}\varphi (t_i),\varphi (t_i).\hfill \end{array}$$ (6.16) Note that by Equation (6.5) for each finite $`k`$, all the operators in Equation (6.16) are of finite $`H`$-rank individually and therefore we can write $$\begin{array}{cc}\hfill \mathrm{tr}_H(Q_{\mathrm{\Delta },k})\mathrm{tr}_H(Q_{\mathrm{\Omega },k})& =\underset{tT}{}\underset{i=1}{\overset{n}{}}Q_{\mathrm{\Delta },k}t_i,t_iQ_{\mathrm{\Omega },k}\varphi (t_i),\varphi (t_i)\hfill \\ & \stackrel{(\text{6.16})}{=}\mathrm{tr}_H(P_{\mathrm{\Delta },k})\mathrm{tr}_H(P_{\mathrm{\Omega },k})\stackrel{(\text{6.5})}{}\frac{1}{L}.\hfill \end{array}$$ (6.17) This implies, since $`\frac{1}{L}`$ is discrete, $$\mathrm{tr}_H(Q_\mathrm{\Delta })\mathrm{tr}_H(Q_\mathrm{\Omega })\stackrel{\text{(}\text{6.12}\text{), (}\text{6.13}\text{)}}{=}\underset{k\mathrm{}}{lim}\left(\mathrm{tr}_H(Q_{\mathrm{\Delta },k})\mathrm{tr}_H(Q_{\mathrm{\Omega },k})\right)\stackrel{(\text{6.17})}{}\frac{1}{L}.$$ (6.18) Putting everything together we see $$\begin{array}{cc}\hfill rdim_H(\mathrm{im}(A))& \stackrel{(\text{6.14})}{=}\underset{tT}{}\underset{i=1}{\overset{n}{}}(P_\mathrm{\Delta }t_i,t_i_{l^2(\mathrm{\Delta })^n}P_\mathrm{\Omega }\varphi (t_i),\varphi (t_i)_{l^2(\mathrm{\Omega })^n})\hfill \\ & \stackrel{(\text{6.15})}{=}\underset{tT}{}\underset{i=1}{\overset{n}{}}(Q_\mathrm{\Delta }t_i,t_iQ_\mathrm{\Omega }\varphi (t_i),\varphi (t_i))\hfill \\ & =\mathrm{tr}_H(Q_\mathrm{\Delta })\mathrm{tr}_H(Q_\mathrm{\Omega })\stackrel{(\text{6.18})}{}\frac{1}{L}.\hfill \end{array}$$ Since $`AM_n(G)`$ was arbitrary, Proposition 6.1 follows. ∎ ###### 6.19 Definition. A group $`G`$ is called treelike (of finity $`r`$), if there are free $`G`$-sets $`\mathrm{\Delta }`$ and $`\mathrm{\Omega }^{}`$, a non-empty trivial $`G`$-set $`X`$ with $`0r`$ elements and an almost $`G`$-map $$\alpha :\mathrm{\Delta }\mathrm{\Omega }^{}X.$$ Julg and Valette show that free groups are treelike of finity $`1`$. On the other hand, Dicks and Kropholler prove that if $`G`$ is treelike of finity $`1`$ then $`G`$ is free. As an immediate consequence of Proposition 6.1 with $`H=\{1\}`$ we get: ###### 6.20 Theorem. Let $`Q`$ be a treelike group of finity $`r`$. Then $`Q`$ fulfills the Atiyah conjecture of order $`\frac{1}{r}`$ over $`Q`$. We generalize in this section the main result in and show that the class of treelike groups of finity $`r`$ coincides with the class of groups which contain a free group of index $`r`$. So, in view of Lemma 2.2, Theorem 6.20 does not provide new cases of the Atiyah conjecture. ###### 6.21 Theorem. The following statements are equivalent for a group $`G`$: 1. $`G`$ is treelike of finity which divides $`r`$. 2. $`G`$ acts on a tree and all vertex stabilizers have order which divides $`r`$. 3. $`G`$ is the fundamental group of a graph of groups where the order of each vertex group divides $`r`$. 4. $`G`$ contains a free subgroup of finite index $`d`$ which divides $`r`$. 5. $`G`$ acts freely on a union of $`d`$ trees where $`d`$ divides $`r`$. 6. $`G`$ contains a free subgroup of finite index and the order of every finite subgroup of $`G`$ divides $`r`$. All statements which don’t involve treelikeness are well known. Those proofs for which no reference in a standard textbook could be found are included here. ###### Proof. (1) $``$ (2): Suppose $`G`$ is treelike of finity $`r`$. That means we have two free $`G`$-sets $`\mathrm{\Delta }`$ and $`\mathrm{\Omega }^{}`$ and a bijective almost $`G`$-map $`\theta :\mathrm{\Delta }\mathrm{\Omega }:=\mathrm{\Omega }^{}(_{i=1}^r\{\})`$. Let $`V`$ be the set of all maps with same domain and range which coincide with $`\theta `$ outside a finite subset of $`\mathrm{\Delta }`$. By \[9, 2.4\] $`V`$ is the vertex set of a $`G`$-tree with finite edge stabilizers. We want to show that the vertex stabilizers are finite, and their order is divisible by $`r`$. That means we want to show that there is no way to change $`\theta `$ at only finitely many points in the domain to obtain an $`f`$ fixed by $`H<G`$ unless $`H`$ is finite and its order divides $`r`$. Remember that the action of $`G`$ on the maps from $`\mathrm{\Delta }`$ to $`\mathrm{\Omega }`$ is given by conjugation: $`f^g(x)=g^1f(gx)`$. Therefore, $`f`$ being an $`H`$-fixed point is equivalent to $`f`$ being an $`H`$-map. Each map $`fV`$ has an index $`i(f):=_{x\mathrm{\Omega }}(\left|f^1(x)\right|1)`$. This is well defined if the inverse image of almost every point of $`\mathrm{\Omega }`$ consists of one point. In particular, it is well defined for the bijective map $`\theta `$ and $`i(\theta )=0`$. If we change a map at one point (or iteratively at finitely many points), this index remains well defined and unchanged (the inverse image of one point will be smaller by one, whereas the inverse image of another point will be larger by one). In particular, $`i(f)=0`$ for every $`fV`$. Assume that $`fV`$ is an $`H`$-map for $`H<G`$. Then $`f`$ will miss, say, $`k`$ of the $`r`$ $`H`$-fixed points of $`\mathrm{\Omega }`$, and the inverse image of the remaining $`rk`$ of those will consist of unions of free $`H`$-orbits. In addition, the inverse image of each point in a given free $`H`$-orbit of $`\mathrm{\Omega }`$ will be of equal size. If $`H`$ is infinite then $`k=r`$ because no point can have an infinite inverse image (else $`i(f)`$ is not defined). But then $`f`$ is not injective because $`i(f)=0`$. Therefore the inverse image of at least one free $`H`$-orbit in $`\mathrm{\Omega }`$ consists of more than one orbit, which is impossible because $`i(f)`$ is defined. This contradiction shows that no infinite subgroup $`H`$ of $`G`$ fixes a point $`fV`$. If $`H`$ is finite and $`f`$ is an $`H`$-map then the fact that $`\mathrm{\Delta }`$ and $`\mathrm{\Omega }^{}`$ are free $`H`$-sets implies as above $`0=i(f)=r+\left|H\right|N`$, i.e. $`r`$ is divisible by the order of $`H`$. (2) $``$ (3): This is the structure theorem for groups acting on a tree, compare e.g. \[8, I.4.1.\]. (3) $``$ (4) This is proved as \[2, Theorem 8.3\]. The statement follows also easily from the arguments of \[8, IV.1.6\], where is is shown that $`G`$ contains a free normal subgroup of finite index. (4)$``$(5): Let $`F`$ be a free subgroup of index $`d`$ in $`G`$. Choose a free basis $`\{f_i\}`$ of $`F`$. Construct the following graph: the elements of $`G`$ are the vertices, and two vertices $`g`$ and $`h`$ are joined by an edge if and only if $`gf_i=h`$ for a suitable $`f_i`$ in the basis of $`F`$. The graph is a free left $`G`$-set in the obvious way (with transitive action on the vertices). The component of the trivial element is the Cayley graph of $`F`$ with respect to the generators $`\{f_i\}`$ and therefore is a tree \[8, I.8.2\]. The vertex sets of the other components are the translates of $`F`$. Consequently, the graph consists of $`d`$ trees. (5) $``$ (1): A construction of Julg and Valette gives an almost $`G`$-map from the set of edges to the set of vertices united with $`d`$ additional points which are fixed under the $`G`$-action. It is done as follows: choose base points for each of the trees. These base points are mapped to the $`d`$ additional points. Each other vertex is mapped to the first edge of the geodesic starting at this vertex and ending at the basepoint of its component. This geodesic and therefore the edge is unique because we are dealing with trees. Since the $`G`$-action was free, we get a treelike structure for $`G`$ with finity $`d`$. (6) $``$ (2): A group which contains a free subgroup of finite index acts on a tree with finite vertex stabilizers \[8, IV.1.6\]. The stabilizers are subgroups of $`G`$, therefore their order divides $`r`$. (3) $``$ (6): By \[8, IV.1.6\], if $`G`$ is the fundamental group of a graph of groups with finite vertex groups where the order is bounded by $`r`$, then it contains a free subgroup of finite index. Moreover, every finite subgroup of $`G`$ is contained in a conjugate of a vertex group \[8, I.7.11\], i.e. the order of every finite subgroup divides $`r`$. ∎
warning/0001/hep-ph0001044.html
ar5iv
text
# The MSSM at present and future colliders ## 1 Introduction Supersymmetric (SUSY) theories possess very appealing theoretical properties (for a review, see e.g. Ref. ) and can certainly be called the currently best motivated extensions of the Standard Model (SM). Their minimal realization, the Minimal Supersymmetric Standard Model (MSSM), postulates superpartners to the SM fields and requires an enlarged Higgs sector with two Higgs doublets giving rise to five physical Higgs-boson states. While the MSSM is minimal in the sense of its particle content, in its unconstrained form (i.e. without specific assumptions about the SUSY-breaking mechanism) it introduces more than 100 free parameters (masses, mixing angles, etc.) in addition to the SM parameters. If low-energy Supersymmetry turns out to be realized in nature and superpartners will be found at the present or the next generation of colliders, the determination of the MSSM parameters will be a very demanding task, both from the experimental and the theoretical side. A precise determination of the model parameters will not only be important in order to investigate whether the MSSM is consistent with the data, but also to infer possible patterns of the underlying SUSY-breaking mechanism from the spectrum of the SUSY particles. In this context it will be important to take advantage of all possible sources of information, i.e. both from the direct production of SUSY particles and from indirect constraints on the model via precision observables. ## 2 Direct production of SUSY particles A detailed investigation of the production and decay processes of SUSY particles is indispensable for the SUSY searches at present and future colliders, as the main background for SUSY signals will often be SUSY itself. For production processes at hadron colliders QCD corrections are very important. In general they give rise to a considerable enhancement of the production cross sections (for recent reviews, see Ref. ). Complementary to the hadron colliders Tevatron and LHC, where in particular the latter has a large discovery potential for a wide range of SUSY processes, an $`e^+e^{}`$ linear collider provides high-precision information for all kinematically accessible SUSY particles (see Ref. for an overview). In the context of constraining the parameters of the model it can be very useful to take advantage of polarization of the $`e^{}`$ and also the $`e^+`$ beam or to study spin correlations in the production and subsequent decay of SUSY particles (see Ref. and references therein). As an example for the production of scalar top quarks at a linear collider with $`\sqrt{s}=500`$ GeV , Fig. 1 shows the determination of the mass of the lightest scalar top quark and the mixing angle in the $`\stackrel{~}{t}`$ sector from the cross sections with polarization of the $`e^{}`$ beam of $`\pm 0.9`$. As shown in the figure, for an integrated luminosity of $`=500`$ fb<sup>-1</sup> a very precise measurement could be possible. ## 3 The lightest Higgs boson in the MSSM In contrast to the SM, the mass of the lightest $`𝒞𝒫`$-even Higgs boson in the MSSM, $`m_\mathrm{h}`$, is not a free parameter, but is calculable from the other parameters of the model. It is bounded to be smaller than the Z-boson mass at the tree level. This bound, however, is strongly affected by large radiative corrections. The dominant corrections arise from the $`t`$$`\stackrel{~}{t}`$ sector of the MSSM. At two-loop order an upper bound on $`m_\mathrm{h}`$ of about $`m_\mathrm{h}\stackrel{<}{}\mathrm{\hspace{0.33em}135}`$ GeV is obtained . This upper bound on $`m_\mathrm{h}`$ is a definite and robust prediction of the MSSM, which can be tested at the present and the next generation of colliders. By comparing the present experimental limit on $`m_\mathrm{h}`$ from the search at LEP2 with the theoretical result for the upper bound on $`m_\mathrm{h}`$ in the MSSM as a function of $`\mathrm{tan}\beta `$, it is possible to derive constraints on $`\mathrm{tan}\beta `$. For recent analyses in this context, see Ref. . If the lightest $`𝒞𝒫`$-even Higgs boson of the MSSM will be found, its mass will be determined with high precision. The prospective accuracy at the LHC is $`\mathrm{\Delta }m_\mathrm{h}=0.2`$ GeV , while at a future linear collider an accuracy of even $`\mathrm{\Delta }m_\mathrm{h}=0.05`$ GeV could be achievable. ## 4 Precision tests of the MSSM Complementary to the direct production processes of SUSY particles, constraints on the model can also be obtained from the virtual contributions of SUSY particles to SM processes. In global fits to the electroweak data taken at LEP, SLC and the Tevatron the fit quality in the MSSM is similar to the SM case . In the low energy regime rare $`B`$ decays have turned out to be sensitive probes for physics beyond the SM . Of particular importance for deriving indirect constraints on the MSSM are the precision observables $`M_\mathrm{W}`$, $`\mathrm{sin}^2\theta _{\mathrm{eff}}`$, and in the future possibly also $`m_\mathrm{h}`$. In Fig. 2 the SM and the MSSM predictions for $`M_\mathrm{W}`$ and $`\mathrm{sin}^2\theta _{\mathrm{eff}}`$, based on the complete one-loop results and the leading higher-order QCD and electroweak corrections (see Ref. and references therein), are compared with the experimental accuracy obtainable at LEP2, SLC and the Tevatron as well as with prospective future accuracies at the LHC and at a high-luminosity linear collider in a dedicated low-energy run (GigaZ) . The experimental accuracies assumed in Fig. 2 for LEP2/Tevatron, LHC and GigaZ are $`\mathrm{\Delta }M_\mathrm{W}=30`$ MeV, 15 MeV, 6 MeV and $`\mathrm{\Delta }\mathrm{sin}^2\theta _{\mathrm{eff}}=1.8\times 10^4`$, $`1.8\times 10^4`$, $`1\times 10^5`$, respectively. The allowed region of the SM prediction corresponds to varying $`m_\mathrm{h}`$ in the interval $`90\mathrm{GeV}m_\mathrm{h}400\mathrm{GeV}`$ and $`m_\mathrm{t}`$ within its present experimental uncertainty, while in the region of the MSSM prediction besides the uncertainty of $`m_\mathrm{t}`$ also the SUSY parameters are varied. As can be seen in the figure, the precision observables $`M_\mathrm{W}`$ and $`\mathrm{sin}^2\theta _{\mathrm{eff}}`$ provide a very sensitive test of the theory, in particular in the case of the GigaZ accuracy. It should be noted that with a future detection of the Higgs boson, a prospective reduction on the experimental error of $`m_\mathrm{t}`$ to $`\mathrm{\Delta }m_\mathrm{t}=2`$ GeV at the LHC and $`\mathrm{\Delta }m_\mathrm{t}=0.2`$ GeV at a linear collider , and with the possible detection of SUSY particles the allowed range of the theory prediction in Fig. 2 will be drastically reduced. The prediction for $`m_\mathrm{h}`$ within the MSSM is particularly sensitive to the parameters in the $`t`$$`\stackrel{~}{t}`$ sector, while in the region of large $`M_\mathrm{A}`$ and large $`\mathrm{tan}\beta `$ (giving rise to Higgs masses beyond the reach of LEP2) the dependence on the latter two parameters is relatively mild. A precise measurement of $`m_\mathrm{h}`$ can thus be used to constrain the parameters in the $`t`$$`\stackrel{~}{t}`$ sector of the MSSM. In Fig. 3 it is assumed that the mass of the lightest scalar top quark, $`m_{\stackrel{~}{\mathrm{t}}_1}`$, is known with high precision, while the mass of the heavier scalar top quark, $`m_{\stackrel{~}{\mathrm{t}}_2}`$, and the mixing angle $`\theta _{\stackrel{~}{\mathrm{t}}}`$ are treated as free parameters. The Higgs boson mass is assumed to be known with an experimental precision of $`\pm 0.5`$ GeV (and a hypothetical value for the central value of $`m_\mathrm{h}`$ is considered), and $`\mathrm{\Delta }m_\mathrm{t}=0.2`$ GeV is used. The figure shows that the values of $`m_{\stackrel{~}{\mathrm{t}}_2}`$, $`\theta _{\stackrel{~}{\mathrm{t}}}`$ which are compatible with a Higgs-mass prediction of $`m_\mathrm{h}=120.5\pm 0.5`$ GeV are given by two narrow bands in the $`m_{\stackrel{~}{\mathrm{t}}_2}`$$`\theta _{\stackrel{~}{\mathrm{t}}}`$ plane (the bands corresponding to smaller and larger values of $`m_{\stackrel{~}{\mathrm{t}}_2}`$ are related to smaller and larger values of the off-diagonal entry in the scalar top mixing matrix, respectively). The uncertainty of $`\mathrm{\Delta }m_\mathrm{t}=0.2`$ GeV assumed in Fig. 3 is seen to have only a marginal effect. Combining the constraints on the parameters in the scalar top sector with the constraints from the precision observables $`M_\mathrm{W}`$ and $`\mathrm{sin}^2\theta _{\mathrm{eff}}`$ and with the information from the direct production of the scalar top quarks (see Fig. 1) will clearly lead to a very sensitive test of the MSSM. In Fig. 3 the theoretical uncertainty in the Higgs-mass prediction from unknown higher-order contributions and the parametric uncertainty related to all parameters besides $`m_{\stackrel{~}{\mathrm{t}}_2}`$, $`\theta _{\stackrel{~}{\mathrm{t}}}`$, and $`m_\mathrm{t}`$ has been neglected. In a more realistic analysis these uncertainties, in particular the dependence on the other SUSY parameters according to the available experimental information on these parameters, will have to be taken into account. The author thanks S. Heinemeyer, W. Hollik and P.M. Zerwas for collaboration and W. de Boer, J. Erler, M. Krämer, S. Kraml, G. Moortgat-Pick, C. Schappacher, A. Sopczak and U. Schwickerath for useful communications.
warning/0001/gr-qc0001030.html
ar5iv
text
# TIME DISPERSION AND EFFICIENCY OF DETECTION FOR SIGNALS IN GRAVITATIONAL WAVE EXPERIMENTS ## 1 Introduction There are today five detectors of gravitational waves (GW) in operation , all of them of the resonant type. It is thus important to study in detail the problem of the coincidence search. In the past, after the initial works of Weber, three papers on coincidence search have been published . These coincidence search was made under two hidden assumptions: a)the signal-to-noise ratio (SNR) was considered to be very large, b)the event time was considered to be equal to the signal time. Since we expect very tiny signals, the study of the problem when dealing with small SNRs is fundamental. This is our object here using simulated signals but with real noise measured with the EXPLORER and NAUTILUS detectors. ## 2 Signal and events In order to clarify the distinction between $`signal`$ and $`event`$ let us recall how an $`event`$ is defined. We describe the procedure adopted by the Rome group, but a similar procedure is adopted also by the ALLEGRO, AURIGA and NIOBE groups. For NAUTILUS and EXPLORER the data have a sampling time of 4.544 ms and are filtered with a filter matched to short bursts for the detection of delta-like signals. The filter makes use of power spectra obtained during periods of two hours. Be $`x(t)`$ the filtered output of the electromechanical transducer which converts the mechanical vibrations of the bar in electrical signals. This quantity is normalized, using the detector calibration, such that its square gives the energy innovation $`E_f`$ for each sample, expressed in kelvin units. In absence of signals, for well behaved noise due only to the thermal motion of the bar and to the electronic noise of the amplifier, the distribution of $`x(t)`$ is normal with zero mean. The variance (average value of the square of $`x(t)`$) is called $`effectivetemperature`$ and is indicated with $`T_{eff}`$. The distribution of $`x(t)`$ is $$f(x)=\frac{1}{\sqrt{2\pi T_{eff}}}e^{\frac{x^2}{2T_{eff}}}$$ (1) After the filtering of the raw-data, $`events`$ are extracted as follows. A threshold is set in terms of a critical ratio defined by $$CR=\frac{|x|<|x|>}{\sigma (|x|)}=\frac{\sqrt{SNR_f}\sqrt{\frac{2}{\pi }}}{\sqrt{1\frac{2}{\pi }}}$$ (2) where $`\sigma (|x|)`$ is the standard deviation of $`|x|`$ and $$SNR_f=\frac{E_f}{T_{eff}}$$ (3) $`T_{eff}`$ is determined by taking the average of the filtered data during the ten minutes preceeding each considered event. The threshold is set at CR=6, in order to have about one or two hundred events per day. This corresponds to an energy $`E_t=19.5T_{eff}`$. When the filtered data go above this threshold, the time behaviour is considered until the filtered data go below the threshold for more than ten seconds. The maximum amplitude and its occurrence time define the $`event`$. By the word $`signal`$ here we mean the response of the detector to an external excitation in absence of noise. It is then evident that an $`event`$ is a combination of signal and noise. In the following we shall use SNR to indicate the ratio between the $`signal`$ energy, which we denote with $`E_s`$ and the noise $`T_{eff}`$, $$SNR=\frac{E_s}{T_{eff}}$$ (4) The effect of the noise on the signal has been discussed in and it turns out to be larger that one could erroneously think. For example, with SNR=20 (for NAUTILUS), one could think that most of the signals would be detected above the threshold $`E_t=19.5T_{eff}`$. It turns out that the detection efficiency is of the order of 50%, as the noise might be in phase with the signal, pushing it even higher over the threshold or in counter-phase, pushing it below the threshold. This means that the detection efficiency for $`m^{pl}`$ coincidences with $`m`$ detectors, in the case $`E_sE_f`$, is of the order of $`\frac{1}{2^m}`$. The noise acts also in producing an $`eventtime`$ different from the time the $`signal`$ was applied. This influences the choice of the coincidence time window. ## 3 Experimental data We use two sets of experimental data, obtained with EXPLORER in 1991 and with NAUTILUS in 1998. This is because the two detectors had their best performance respectively in 1991 and 1998, and also because their detection bandwidth is very different in the two cases. The main characteristics of these two detectors are given in table 1. The Q value for NAUTILUS is small because of electrical losses in the transducer. Work is in progress to obtain a larger Q value. Both EXPLORER and NAUTILUS are equipped with similar resonant capacitive transducers, thus they have two resonance modes at frequencies of 904.7 Hz and 921.3 Hz for EXPLORER and 907.0 Hz and 922.5 Hz for NAUTILUS. The algorithm for extracting small delta signals from the noise is based on the measurement of the power spectra and it takes care of both resonance modes . Applying a delta signal to the detector we have at the transducer output the sum of the two mode oscillations, sharply beginning at the time the pulse was applied and decaying with a time constant proportional to the Q value. The filter operates a sort of weighted average and the result V(t) has maximum value at the time the delta was applied (t=0) and oscillates, with envelope obeying the equation $$V(t)=V_oe^{\beta _3|t|}$$ (5) The quantity $`\beta _3`$ divided by $`\pi `$ gives the frequency bandwidth of the apparatus. An example of the behaviour of the filtered signal with time and in absence of noise is shown in fig.1 for the two detectors. For the filtered data we get $`\beta _3=6.0\frac{rad}{s}`$ for EXPLORER and $`\beta _3=0.39\frac{rad}{s}`$ for NAUTILUS. The bandwidth of EXPLORER in 1991 was then $`\mathrm{}f=\frac{\beta _3}{\pi }=1.9Hz`$, the bandwidth for NAUTILUS in 1998 was $`\mathrm{}f=\frac{\beta _3}{\pi }=0.12Hz`$. This small bandwidth will be increased in future with improved transducers and electronics . ## 4 Simulation with delta signals We make use of eight hours of data recorded with NAUTILUS on 12 July 1998 ($`T_{eff}=4.18mK)`$ and we use four hours for EXPLORER recorded on 13 September 1991 ($`T_{eff}=6.08mK)`$. In absence of applied signals 34 events are detected for EXPLORER and 41 events for NAUTILUS, due to the noise fluctuation. These events are vetoed in all the successive analyses made with applied signals. Delta signals with given $`SNR`$ are applied over the real noise with a certain periodicity. One must make sure that the filtering of a new applied signal is not disturbed by the residual of the previous applied signal. This is obtained if the periodicity of the applied signals is much larger than $`\frac{1}{\beta _3}`$. Thus we have used for EXPLORER a periodicity of half a minute for large SNR and a periodicity of ten seconds for smaller SNR. For NAUTILUS the periodicities are one minute and twenty seconds. The signals are applied at the exact time the data are sampled with a sampling rate of 4.544 ms. For EXPLORER we have found the result given in table 2. For NAUTILUS we have found the result given in table 3. The efficiency is also shown in fig.2 The theoretical probability to detect a signal with a given SNR, in presence of a well behaved Gaussian noise, is calculated as follows. We put $`y=(s+x)^2`$ where $`s\sqrt{SNR}`$ is the signal we look for and $`x`$ is the gaussian noise. We obtain easily $$probability(SNR)=_{SNR_t}^{\mathrm{}}\frac{1}{\sqrt{2\pi y}}e^{\frac{(SNR+y)}{2}}cosh(\sqrt{ySNR})𝑑y$$ (6) where we put $`SNR_t=19.5`$ for the present EXPLORER and NAUTILUS detectors. The theoretical efficiency as deduced from eq.6 is reported in tables 2 and 3 and in fig.2. We notice a deviation between experimental and theoretical efficiencies at small SNR. This is due to the non gaussian character of the real noise. The time when the event due to a signal is observed deviates from the time the signal is applied. We show in fig. 3 the standard deviation against SNR for EXPLORER 1991 and for NAUTILUS 1998. The lines are the best fits with the following equations: EXPLORER $`\frac{1}{2\pi 1.74\pm 0.08}\sqrt{\frac{2}{SNR}}`$ NAUTILUS $`\frac{1}{2\pi 0.124\pm 0.007}\sqrt{\frac{2}{SNR}}`$. We can write the empirical formula $$\sigma =\frac{1}{2\pi \mathrm{\Delta }f}\sqrt{\frac{2}{SNR}}$$ (7) We see, as expected, that the time deviation decreases linearly with increasing bandwidth. If we extrapolate to a SNR=100 and with a target bandwidth for resonant detectors of the order of $`\mathrm{\Delta }f50Hz`$ we find a possible time resolution of the order of less than one millisecond, as already recognized with room temperature experiments . The delay distributions for signals with SNR=30 and SNR=10 are shown in fig. 4. We note for NAUTILUS a few events with delay greater than 1 s with respect to the time of the applied signals. We have asked ourselves how it is possible to have a time deviation over 1 s for signals with SNR=30. This is due to the fact that the noise, although the data were selected so to have small $`T_{eff}5mK`$, has not completely a gaussian character. We have considered the particular case of the event (fig.4, SNR=30) detected with the NAUTILUS data 1.422 s $`before`$ the signal was applied. In order to understand this result we plot in fig.5 the behaviour of the signal with zero noise, of the noise alone and of the signal added to the noise. For this particular case if we raise the signal to SNR=50 the corresponding event has a time delay of -64 ms (still not quite zero). If an additional filter is applied to the data, such as to require i.e. that the detected event behaved in a gaussian way, the signal with SNR=30 is lost, in spite of being a $`delta`$ signal. We remark also that the events have energy different from that of the signal. This is shown in tables 2,3 and in fig.6 where we give the distributions of the ratio $`\frac{E_f}{E_s}`$ for SNR=30 and SNR=10. Finally we make the following consideration for the case when multiple coincidences with $`M`$ detectors are searched for. From tables 2 and 3 we deduce that when the signal is near the threshold the efficiency of detection in nearly 50%. This means that for these signals the total efficiency for $`M^{pl}`$ coincidences is $`\frac{1}{2}^m`$. Since we want an efficiency near unity (because of the very few possible GW signals) we must consider only signals with $`SNR`$ at least twice the value $`SNR_t`$ of the threshold. ## 5 Conclusions We have studied the $`events`$ generated in a resonant GW detector when excited by GW bursts with $`SNR`$ near the threshold $`SNR_t`$ used for defining the events. For $`SNR=SNR_t`$ the detection efficiency is nearly $`\frac{1}{2}`$. The efficiency goes to 100% for $`SNR>2SNR_t`$, and it is still $`>10\%`$ for $`SNR\frac{SNR_t}{2}`$. The time of the event might be different from that of the signal, with standard deviation depending on the SNR and on the bandwidth of the experimental apparatus. In this analysis we have applied delta signals at the exact time of the samples. If the delta signals are applied randomly, as in the real case, the efficiency will be smaller and the time dispersion larger. Delta-like signals can be lost if the requirement to satisfy the theoretical behaviour expected for a delta signal is imposed on the detected events, even for $`SNR=30`$, as shown in fig.5. This can jeopardize a search looking for very rare gravitational wave signals.
warning/0001/hep-ex0001048.html
ar5iv
text
# Decay ϕ→𝜋⁺⁢𝜋⁻ ## 1 Introduction The decay $`\varphi \pi ^+\pi ^{}`$ reveals itself as an interference pattern in the energy dependence of the cross section of the process $`e^+e^{}\pi ^+\pi ^{}`$ in the region close to $`\varphi `$ peak. The $`\varphi \pi ^+\pi ^{}`$ decay was previously studied at VEPP-2M collider and current PDG value $`B(\varphi \pi ^+\pi ^{})=(8_4^{+5})10^5`$ is based on these results. The decay $`\varphi \pi ^+\pi ^{}`$ violates both OZI rule and G-parity conservation. The decay amplitude in Vector Dominance Model (VDM) was calculated in . The main contribution into the amplitude of the $`\varphi \pi ^+\pi ^{}`$ decay in this work comes from the electromagnetic $`\varphi \rho `$ mixing. The contribution of the $`\varphi \rho `$ transitions through the $`\omega `$ meson and other intermediate states such as $`K\overline{K}`$, $`\eta \gamma `$, etc. is estimated to be $`20\%`$ of electromagnetic $`\varphi \rho `$ mixing. The value of the branching ratio of the decay $`\varphi \pi ^+\pi ^{}`$ calculated from the decay amplitude obtained in the work is almost 2 times higher than current PDG value . Different $`\varphi \omega `$ mixing models were scrutinized in respect to this decay in . The branching ratio calculated in this work is lower than that in , but discrepancy between the experimental results, especially , and the theoretical prediction still exists. Possible mechanisms, which could decrease the theoretical branching ratio, are discussed in . One of them is the existence of direct decay $`\varphi \pi ^+\pi ^{}`$. ## 2 Experiment The experiments with SND detector (Fig. 1) at VEPP-2M $`e^+e^{}`$ collider are being conducted since 1995. SND is a general purpose non-magnetic detector . The main part of the SND is a 3-layer spherical electromagnetic calorimeter, consisting of 1632 NaI(Tl) crystals . The solid angle of the calorimeter is $`90\%`$ of $`4\pi `$ steradian. The angles of charged particles are measured by two cylindrical drift chambers covering 95% of full solid angle. The important part of the detector for the process under study is the outer muon system, consisting of streamer tubes and plastic scintillation counters. The 1998 experiment was carried out in the energy range $`2\mathrm{E}_b`$=984–1060 MeV in 16 energy points and consisted of 2 data taking runs : PHI\_9801, PHI\_9802. The total integrated luminosity $`\mathrm{\Delta }L=8.6`$ pb<sup>-1</sup> collected in these runs corresponds to $`13.210^6`$ produced $`\varphi `$ mesons. The integrated luminosity was measured using $`e^+e^{}e^+e^{}`$ events selected in the same acceptance angle as the events of the process under study. The interference term in the $`e^+e^{}e^+e^{}`$ cross section due to $`\varphi e^+e^{}`$ decay was also taken into account. The systematic error of the luminosity measurement was estimated to be 2%. ## 3 Event selection The energy dependence of the cross section of the process $$e^+e^{}\pi ^+\pi ^{}$$ (1) was studied in the vicinity of $`\varphi `$ meson. Events containing two collinear charged particles and no photons were selected for analysis. The following cuts on angles of acollinearity of the charged particles in azimuthal and polar directions were imposed: $`\mathrm{\Delta }\phi <10^{}`$, $`\mathrm{\Delta }\theta <25^{}`$. To suppress the beam background the production point of charged particles was required to be within $`0.5`$ cm from the interaction point in the azimuthal plane and $`\pm 7.5`$ cm along the beam direction (the longitudinal size of the interaction region $`\sigma _z`$ is about $`2`$ cm). The polar angles of the charged particles were required to be in the range $`45^{}<\theta <135^{}`$, determined by acceptance angle of the muon system. The main sources of background are cosmic muons and the following processes: $`e^+e^{}e^+e^{},`$ (2) $`e^+e^{}\mu ^+\mu ^{},`$ (3) $`e^+e^{}\pi ^+\pi ^{}\pi ^0,`$ (4) $`e^+e^{}K_SK_L.`$ (5) To suppress the background from the process (2) a procedure of $`e/\pi `$ separation was used. It utilizes the difference in the longitudinal energy deposition profiles in the calorimeter for electrons and pions. The separation parameter was calculated for each charged particle in an event: $$K=\mathrm{log}\left(\frac{𝒫_e(E_1,E_2,E_3,E_e)}{𝒫_\pi (E_1,E_2,E_3,E_\pi )}\right),$$ (6) where $`𝒫_{e(\pi )}`$ — the probability for an electron (pion) with the energy $`E_{e(\pi )}`$ to deposit the energy $`E_i`$ in the $`i`$-th calorimeter layer. $`E_{e(\pi )}`$ in our case is equal to the beam energy. The separation parameters distribution for both particles in collinear events with no hits in the muon system, is shown in Fig. 2. This distribution is asymmetric because the particles are ordered according to their energy depositions in the calorimeter. To select the events of the process (1) the cut $`K_1+K_2<0`$ was imposed. The background from the process (2) was suppressed by a factor of $`3000`$, while only 7% of the events of the process under study were lost. Remaining background from the process $`e^+e^{}e^+e^{}`$ was about 1.5%. The events of the process (3) and cosmic muons can be efficiently suppressed by the muon system. We required no hits in the scintillation counters of the muon system. The efficiency of these counters was estimated using cosmic muons selected by special cuts. Due to possible admixture of beam events which actually produce no hits in the muon counters only the lower boundary of the efficiency was obtained: 99.8%. Thus estimated contribution of cosmic muons does not exceed 0.7% of the total number of events of the process (1) and was neglected. The energy dependences of the probabilities for muons and pions to produce hits in outer scintillation counters were obtained from the experimental data. With energy increasing from 492 MeV up to 530 MeV these probabilities rise from 84% up to 94% for muons and from 0.5% to 11% for pions. In the final selection of the process (1) the background from the process (3) was about 15%. To suppress the resonant background from the processes (4) and (5) the following cuts on energy depositions in the calorimeter were applied: 1. the energy deposition in the first calorimeter layer of the most energetic particle in an event is less than 75 MeV; 2. the energy deposition in the third calorimeter layer of the least energetic particle in an event is more than 50 MeV. In the events of the process (4), which satisfy the geometrical cuts, the energetic photon from $`\pi ^0`$ decay propagates along the direction of a charged pion producing unusually large energy deposition in the first calorimeter layer for this pion. Such events are suppressed by the first cut. The pions from process (5) are relatively soft with a maximum energy of about 300 MeV and low probability of significant energy deposition in the third calorimeter layer. The second cut is crucial for the rejection of the process (5). The residual cross sections of resonant background processes were estimated by Monte Carlo (MC) simulation: 0.06 nb for the process (4), 0.09 nb for the process (5). To determine the remaining resonant background more accurately the selected events were divided into two data samples using the parameter $`\mathrm{\Delta }\phi `$: $`\mathrm{\Delta }\phi <5^{}`$ and $`\mathrm{\Delta }\phi >5^{}`$. The resolution in $`\mathrm{\Delta }\phi `$ is about $`1^{}`$. The main part of the events of the process (1) is contained in the first sample. Due to the emission of hard photons by initial or final particles and errors in the reconstruction of the particle angles some events of the process (1) can migrate into the second sample. The level of the resonant cross section $`\sigma _2^{res}`$, determined in the second sample, was used to estimate the resonant background in the first sample: $`\sigma _1^{res}=k\sigma _2^{res}`$. The coefficient $`k=1.5\pm 0.3`$ was obtained by MC simulation of the processes (4) and (5), its error is determined by accuracy of simulation of energy depositions of pions in the calorimeter. Because the level of the resonant background is low, the error in $`k`$ does not give significant contribution into the errors of the interference parameters. The cut $`\mathrm{\Delta }\phi <5^{}`$ reduces the level of resonant background down to as low as 0.09 nb. This value is less than 1% of the process (1) detection cross section. The pion polar angle distribution for the process (1) at beam energy higher than 520 MeV is shown in Fig. 3. At this energy the cross sections of the resonant processes (4) and (5) are small. The additional cut on the total energy deposition in the calorimeter $`E_{tot}>400`$ MeV rejects the events of the process (3). A good agreement between experimental distribution and the simulation of the process (1) shows that selected pion sample is quite pure and the level of QED background is low. ## 4 Data analysis The fitting of the detection cross sections for the first and the second samples were performed simultaneously (Fig. 45). To describe the cross sections the following formulae were used: $`\sigma _1^{vis}(E)`$ $`=`$ $`\sigma _{\pi \pi }^{vis}(E)+\sigma _{\mu \mu }^{vis}(E)`$ (7) $`+\sigma _{ee}^{vis}(E)+\sigma _1^{res}(E),`$ $`\sigma _2^{vis}(E)`$ $`=`$ $`C+D(Em_\varphi )`$ (8) $`+\sigma _2^{res}(E),`$ $`\sigma _1^{res}(E)`$ $`=`$ $`k\sigma _2^{res}(E),`$ $`\sigma _2^{res}(E)`$ $`=`$ $`\epsilon _{res}(0.39\sigma _{\pi ^+\pi ^{}\pi ^0}(E)`$ $`+0.61\sigma _{K_SK_L}(E)),`$ where $`E`$ is the CM energy; $`\sigma _{\pi \pi }^{vis}(E)`$ — the detection cross section of the process (1); $`\sigma _{\mu \mu }^{vis}(E)`$ — the contribution of the process (3) (this process was studied in our work ); $`\sigma _{ee}^{vis}(E)=0.2(nb)(m_\varphi /E)^2`$ — the contribution of the process (2). The ratio 0.39:0.61 between the processes (4) and (5) was taken from the simulation. The coefficients $`C`$, $`D`$ and $`\epsilon _{res}`$ were free fit parameters. The following expression was used for $`\sigma _{\pi \pi }^{vis}`$: $`\sigma _{\pi \pi }^{vis}(E)`$ $`=`$ $`\sigma _0(E)R(E)\left|1Z_\pi {\displaystyle \frac{m_\varphi \mathrm{\Gamma }_\varphi }{\mathrm{\Delta }_\varphi (E)}}\right|^2,`$ $`\sigma _0(E)`$ $`=`$ $`{\displaystyle \frac{\pi \alpha ^2\beta ^3(E)F_\pi (E)^2}{3E^2}},`$ (9) where $`\alpha `$ is the fine structure constant; $`\beta (E)=(14m_\pi ^2/E^2)^{1/2}`$; $`m_\varphi `$, $`\mathrm{\Gamma }_\varphi `$, $`\mathrm{\Delta }_\varphi (E)=m_\varphi ^2E^2iE\mathrm{\Gamma }(E)`$$`\varphi `$-meson mass, width and propagator respectively; $`\sigma _0(E)`$ — the Born cross section of the process $`e^+e^{}\pi ^+\pi ^{}`$; $`Z_\pi `$ — complex parameter characterizing strength of the interference. Two representations of $`Z_\pi `$ are used in different works: $`Z_\pi =Q_\pi e^{i\psi _\pi }=\mathrm{𝐑𝐞}Z_\pi +i\mathrm{𝐈𝐦}Z_\pi `$. $`F_\pi (E)`$ is the pion form factor without $`\varphi `$-meson contribution: $$F_\pi (E)^2=F_\pi ^\varphi ^2(1+A(Em_\varphi )+B(Em_\varphi )^2),$$ (10) with $`F_\pi ^\varphi `$ as the pion form factor at the maximum of $`\varphi `$ resonance. $`Q_\pi `$, $`\psi _\pi `$, $`A`$, $`B`$ and $`F_\pi ^\varphi ^2`$ are free fitting parameters. $`R(E)`$ is a factor taking into account detection efficiency and radiative corrections: $$R(E)=\epsilon _\pi \frac{\sigma _{\pi \pi }(E)}{\sigma _0^{}(E)\left|1Q_\pi ^{}e^{i\psi _\pi ^{}}\frac{m_\varphi \mathrm{\Gamma }_\varphi }{\mathrm{\Delta }_\varphi (E)}\right|^2}.$$ (11) $`\sigma _{\pi \pi }`$ is the result of MC integration of differential cross section of the process (1) with all geometrical restrictions . Since the probability for pions to hit the outer scintillation counters depends on energy, it was taken into account during $`\sigma _{\pi \pi }`$ calculation. The remaining contributions into the detection efficiency do not depend on CM energy and pions energies and were included into $`\epsilon _\pi `$. The value $`\epsilon _\pi =0.234`$ was obtained using MC simulation and experimental data. It is mainly determined by the cuts on energy depositions. Its independence of the pions energy was checked in the range 430 – 530 MeV using the pions from the process (4) with energies up to 450 MeV and pions from the process (1) at the beam energy 530 MeV. The geometrical cuts and the requirement on no hits in the outer scintillation counters led to 50% efficiency losses, so the total detection efficiency of the process (1) was approximately 12% at $`E=m_\varphi `$. The $`R(E)`$ was calculated by iteration method. As a first approximation the interference parameters $`Q_\pi ^{}`$ and $`\psi _\pi ^{}`$ from were used. The pion form factor was taken from while calculating $`\sigma _0^{}(E)`$. After fitting $`R(E)`$ was recalculated with corrected $`Q_\pi ^{}`$ and $`\psi _\pi ^{}`$. This procedure was repeated until convergence was reached. The branching ratio $`B(\varphi \pi ^+\pi ^{})`$ is related to the interference parameters by the following formula: $$B(\varphi \pi ^+\pi ^{})=\frac{Q_\pi ^2\alpha ^2\beta ^3(m_\varphi )F_\pi ^\varphi ^2}{36B(\varphi e^+e^{})},$$ (12) where $`B(\varphi e^+e^{})=(2.99\pm 0.08)10^4`$ . The fitting has been performed for each experimental run separately. The results are listed in Table 1. The fit parameters for two runs are in statistical agreement, therefore combined fit was performed to obtain the final results also listed in Table 1. The observed level of resonant background 0.07 nb is in a good agreement with the MC estimation of 0.09 nb. The fitted values of the coefficients $`A`$ and $`B`$ from the equation (10) are $`A=(8.5\pm 0.3)10^3MeV^1`$ and $`B=(4.9\pm 1.0)10^5MeV^2`$. To check the accuracy of the process (3) background subtraction, the fit to the data with more stringent event selection cuts has been done. The additional requirement that the total energy deposition in the calorimeter is higher than 400 MeV significantly reduced the muon background. The obtained interference parameters: $`Q_\pi =0.073\pm 0.006,`$ $`\psi _\pi =(32\pm 5)^{}`$ agrees well with the results from the Table 1. The additional contribution into the shape of interference pattern may come from the process $$e^+e^{}\varphi f_0\gamma \pi ^+\pi ^{}\gamma .$$ (13) The process (13) interferes with the process (1) when soft photon is emitted by pions. This contribution estimated using CMD-2 analysis of the process $`e^+e^{}\pi ^+\pi ^{}\gamma `$ in the vicinity of $`\varphi `$ resonance does not exceed 1.5% of the interference under study. This value was included into the systematic error. The representation $`Z_\pi =\mathrm{𝐑𝐞}Z_\pi +i\mathrm{𝐈𝐦}Z_\pi `$ is suitable to present the different contributions into the systematic error of the interference parameters: 1. the calculation of the radiative corrections: $`\mathrm{𝐑𝐞}Z_\pi `$ — 5%, $`\mathrm{𝐈𝐦}Z_\pi `$ — 3%; 2. the subtraction of the non-resonant background: $`\mathrm{𝐑𝐞}Z_\pi `$ — 0.8%, $`\mathrm{𝐈𝐦}Z_\pi `$ — 0.6%; 3. the contribution of the process (13): $`\mathrm{𝐑𝐞}Z_\pi `$ — 1.5%, $`\mathrm{𝐈𝐦}Z_\pi `$ — 1.5%; 4. the model dependence on the choice of the function approximating the pion form factor: $`\mathrm{𝐑𝐞}Z_\pi `$ — 1%, $`\mathrm{𝐈𝐦}Z_\pi `$ — 8%; 5. the subtraction of the resonant background: $`\mathrm{𝐈𝐦}Z_\pi `$ — 3%. The systematic error of $`F_\pi ^\varphi ^2`$ is determined by the error of the detection efficiency $`\epsilon _\pi `$ ($``$ 5%) and the accuracy of luminosity determination (2%). The final results are the following: $`F_\pi ^\varphi ^2`$ $`=`$ $`2.98\pm 0.02\pm 0.16,`$ (14) $`Q_\pi `$ $`=`$ $`0.073\pm 0.005\pm 0.004,`$ $`\psi _\pi `$ $`=`$ $`(34\pm 4\pm 3)^{},`$ $`B(\varphi \pi ^+\pi ^{})`$ $`=`$ $`(7.1\pm 1.1\pm 0.9)10^5.`$ For another representation of $`Z_\pi `$ we obtained: $`\mathrm{𝐑𝐞}Z_\pi `$ $`=`$ $`0.061\pm 0.005\pm 0.003,`$ $`\mathrm{𝐈𝐦}Z_\pi `$ $`=`$ $`0.041\pm 0.006\pm 0.004.`$ ## 5 Discussion The obtained value of the branching ratio $$B(\varphi \pi ^+\pi ^{})=(7.1\pm 1.1\pm 0.9)10^5$$ agrees well with the world average value $`B(\varphi \pi ^+\pi ^{})=(8_4^{+5})10^5`$ and has a 3 times higher accuracy. However there is a discrepancy between our result and the preliminary result of CMD-2 experiment : $`B(\varphi \pi ^+\pi ^{})=(18.1\pm 2.5\pm 1.9)10^5.`$ The measured value $`\mathrm{𝐈𝐦}Z_\pi =0.041\pm 0.006\pm 0.004`$ agrees with the theoretical predictions while the value $`\mathrm{𝐑𝐞}Z_\pi =0.061\pm 0.005\pm 0.003`$ is 2.5 times lower than the expected value. The different models of the $`\varphi \omega `$ mixing were examined in the work . The lowest value $`\mathrm{𝐑𝐞}Z_\pi ^{th}=0.12`$ from this work also contradicts our results. This disagreement could be understood if the direct decay $`\varphi \pi ^+\pi ^{}`$ exists or/and in case of nonstandard $`\rho \omega \varphi `$ mixing. One can notice that the measured branching ratio of another rare decay $`\varphi \omega \pi ^0`$ , which violates OZI rule and G-parity, disagrees with theoretical predictions. ## 6 Acknowledgement This work is supported in part by Russian Fund for basic researches (grant 99-02-16815) and STP ”Integration” (No.274).
warning/0001/hep-ph0001335.html
ar5iv
text
# 1 The couplings log|𝑔_{00⁢𝑘}| and log|𝑔_{01⁢𝑘}| (with 𝑔 in units of 𝑀_𝑃⁻¹) as functions of log𝑘, for 𝑅=0.2 mm and 𝐿=1 TeV-1. The sharp dips in the curves are bad renderings of a function with narrowly spaced zeros. ## Acknowledgements We kindly thank E. Alvarez, K. Benakli, A. Cohen, S. Dimopoulos, G. Dvali, F. Feruglio, C. Gómez, L. E. Ibáñez, E. Lopez, W. Lerche, J. March-Russell, C. Muñoz, T. Ortin, R. Sundrum and F. Zwirner for comments and discussions. A. D. acknowledges the I.N.F.N. for financial support. S. R. acknowledges the European Union for financial support through contract ERBFMBICT972474 and the Dept. of Physics of the University of Michigan, Ann Arbor. The work of A. D., M. B. G. and S. R. was partially supported by the CICYT project AEN/97/1678. Note added. Dudas and Mourad and Cullen et al. have recently posted , in which extra-dimensional signatures are studied in a complementary framework –stringy dynamics– that also goes beyond the low-energy effective considerations of .
warning/0001/math0001066.html
ar5iv
text
# On some moment maps and induced Hopf bundles in the quaternionic projective space ## 1. Introduction The main motivation for the present work is to understand some aspects of the Riemannian geometry of the focal set $`Foc_{P^n}P^n`$, i.e. of the set of points in the quaternionic projective space $`P^n`$ that are critical values of the normal exponential map with respect to the totally geodesic submanifold $`P^nP^n`$. Our starting point is the Habilitationsschrift of J. Berndt , , where in particular is proved that $`Foc_{P^n}P^n`$ fibers in circles over $`Gr_2(^{n+1})`$, the Grassmannian of 2-planes in $`^{n+1}`$. In the simplest cases $`n=1,2,`$ these focal points distribute on spheres $`S^1`$, $`S^5,`$ respectively and $`n=3`$ thus seems to be the first significant case. Indeed, since $`Foc_{P^3}P^3`$ is a circle bundle over the Klein quadric $`Gr_2(^4)Q^4P^5`$, both the classical projective geometry of $`Q^4`$ (described for example in , pp. 26-35) and its two quaternion Kähler structures can be related to the total space $`Foc_{P^3}P^3`$. More generally, the twofold Kähler-Einstein and quaternion Kähler properties of the Grassmannian $`Gr_2(^{n+1})`$ suggested to us a link between the focal set $`Foc_{P^n}P^n`$ and the Sasakian-Einstein geometry of the induced Hopf bundles considered in our previous work . This paper begins by giving an explicit identification of the focal set $`Foc_{P^n}P^n`$ with the zero level set of the moment map $`\mu `$ associated to the diagonal $`U(1)`$ \- action on $`P^n`$. This is the simplest isometric action giving examples of reductions in quaternion Kähler geometry, and it is well-known that the reduced manifold is precisely the quaternion Kähler Wolf space $`Gr_2(^{n+1})`$ , . According to the notation in use, we write $`P^n///U(1)=Gr_2(^{n+1})`$ to indicate this reduction procedure; this has been proved to be the unique one with respect to isometric circle actions and positive quaternion Kähler manifolds , . Next, the diagonal $`Sp(1)`$-action on $`P^n`$ has as reduction the quaternion Kähler manifold $`P^n///Sp(1)=\stackrel{~}{G}r_4(^{n+1})`$, the Grassmannian of oriented 4-planes in $`^{n+1}`$. The zero set of the corresponding moment map $`\nu :P^n`$ can also be related to focal sets in $`P^n`$. (For these identifications cf. Theorems 3.1, 3.2). On the other hand, the zero level sets $`\mu ^1(0)`$ and $`\nu ^1(0)`$ of these moment maps $`\mu `$ and $`\nu `$ can be identified with the total spaces of some induced Hopf $`S^1`$-bundles. This enables us to define on them a Sasakian-Einstein and a 3-Sasakian metric, respectively, and hence to find a diagram where these zero sets $`\mu ^1(0)`$ and $`\nu ^1(0)`$ are base spaces of $`S^3`$-bundles, now induced by the Hopf fibration $`S^{4n+3}P^n`$. The corresponding total spaces are the Stiefel manifolds $`V_2(^{n+1})`$ and $`\stackrel{~}{V}_4(^{n+1})`$, respectively. The $`3`$-Sasakian structure of the fibers $`S^3`$ and the Sasakian structure of the base spaces $`\mu ^1(0)`$ and $`\nu ^1(0)`$ allow to construct a complex structure on $`V_2(^{n+1})`$ and on $`\stackrel{~}{V}_4(^{n+1})`$, whose definition is very much in the Calabi-Eckmann spirit. On the former Stiefel manifold, this complex structure is not compatible with its standard hypercomplex structure, obtained in , , . On the other hand, both $`V_2(^{n+1})`$ and $`\stackrel{~}{V}_4(^{n+1})`$ are total spaces of framed bundles in Hopf surfaces over Kähler Einstein manifolds. Then these complex structures can be seen to belong to one-parameter families, also suitable for some exceptional cases, as shown in a forthcoming paper, . Acknowledgements. This paper was written while the first named author was visiting the University La Sapienza of Rome and the Max Planck Institut für Mathematik in Bonn. He thanks both institutions for financial support and for hospitality. The second author acknowledges support from MURST project ”Proprietà geometriche delle varietà reali e complesse”. Both authors thank Victor Vuletescu for some helpful discussions, and John C. Wood who accepted to correct our English and patiently read the whole paper. ## 2. Preliminaries We collect here some definitions and basic facts that will be used throughout the paper. Let $`(M,g)`$ be a Riemannian manifold and let $`N`$ be an isometrically immersed submanifold. The critical values of the restriction $`exp_{T^{}N}`$ of the exponential map of $`M`$ to the normal bundle of $`N`$ are called *focal points of $`N`$* and the set of all focal points, here denoted by $`Foc_MN`$, is called *the focal set of $`N`$* (see for example p. 23, , p. 227, p. 283). The focal set of a submanifold may not be a submanifold: for smooth plane curves and for regular surfaces in $`^3`$ the focal sets are respectively the evolutes and the surfaces of the centres, and both of them can have singular points, cf. , pp. 237-238 and p. 232. However, for the examples considered in the present paper all the focal sets turn out to be smooth. Indeed, it seems that not many focal sets of Riemannian submanifolds have been explicitly determined, but we can mention totally geodesic spheres in spheres , and hypersurfaces in space forms , . We are mainly interested in $`Foc_{P^n}P^n`$, the focal set of the totally geodesic $`P^n`$ in the quaternionic projective space $`P^n`$. This focal set appears in J. Berndt’s work , , denoted there by $`Q^n`$, and studied in connection with both the complex Kähler and the quaternion Kähler structure of $`Gr_2(^{n+1})`$, the Grassmannian of complex two-planes in $`^{n+1}`$. Two geometric structures which appear naturally in our context are the Sasakian and the $`3`$-Sasakian ones. We briefly recall the definitions, referring the reader to the survey for further information. A Sasakian manifold is a $`(2n+1)`$-dimensional Riemannian manifold $`(S,g)`$ equipped with a unitary Killing vector field $`\xi `$ such that the field of endomorphisms $`\phi :=\xi `$ satisfies the differential equation $$\phi =Id\eta g\xi ,$$ where $``$ is the Levi-Civita connection of $`g`$ and $`\eta `$ is the dual 1-form of $`\xi `$. The data of a Sasakian structure on the manifold $`S`$ is easily seen to be equivalent to the requirement that the cone metric $`dr^2+r^2g`$, on $`_+\times S`$ have holonomy contained in $`U(n+1)`$. Note that $`\eta `$ is a contact form on $`S`$, hence $`\xi `$ is its Reeb field. In the simplest example, the Euclidean sphere $`S^{2n+1}`$, the Killing vector field is $`\xi =JU`$, $`J`$ being the standard complex structure of $`^{n+1}`$ and $`U`$ the unit outward normal to the sphere. More generally, we look at the induced Hopf $`S^1`$-bundle $`\pi :VM`$, over a smooth submanifold of $`P^N`$. Its total space $`V`$ carries a Sasakian structure $`(V,g,\xi )`$ induced from the one of $`(S^{2N+1},can)`$. If the Fubini-Study metric of $`P^N`$ induces an Einstein metric $`h`$ with Einstein constant $`\alpha `$ on $`M`$, it can be seen (cf. , lemma 1) that the Ricci tensor of the metric $`g`$ has the form $$Ric(g)=\lambda g+\mu \eta \eta ,$$ with $`\lambda =\alpha 2,`$ $`\mu =dimV+1\alpha `$. This is known as the $`\eta `$-Einstein property in Sasakian geometry and, following S. Tanno , an $`\eta `$-Einstein Sasakian metric can be deformed to a Sasakian-Einstein one by setting (2.1) $$g^{}=Ag+A(A1)\eta \eta ,$$ so that $`\xi ^{}=A^1\xi `$ with $`A=\frac{\lambda +2}{dimV+1}`$. Now $`\pi `$ is a Riemannian submersion with respect to the metrics $`h`$ and $`g`$; thus to have a Riemannian submersion with respect to $`g^{}`$, we must consider the scaled metric $`h^{}=Ah`$ on the base $`M`$. As Sasakian geometry may be viewed as the odd-dimensional counterpart of Kähler geometry, the odd-dimensional counterparts of hyperKähler manifolds are $`3`$-Sasakian ones (cf. ). More precisely, a $`(4n+3)`$-dimensional Riemannian manifold $`(S,g)`$ is said to be $`3`$-Sasakian if it is endowed with three mutually orthogonal unit Killing vector fields $`\xi _1`$, $`\xi _2`$, $`\xi _3`$, each one defining a Sasakian structure and satisfying the conditions: $$[\xi _1,\xi _2]=2\xi _3,[\xi _2,\xi _3]=2\xi _1,[\xi _3,\xi _1]=2\xi _2.$$ As above, an equivalent definition requires that the cone metric $`dr^2+r^2g`$ on $`_+\times S`$ be hyperkähler, i.e. its holonomy group be contained in $`Sp(n+1)`$. $`3`$-Sasakian manifolds are necessarily Einstein with positive scalar curvature and their Einstein constant is $`4n+2`$. Given a positive quaternion Kähler manifold $`P`$, one constructs its Kähler-Einstein twistor space $`Z_P`$ and then an $`S^1`$ bundle over it whose Chern class is, up to torsion, that of an induced Hopf bundle. The total space $`S`$ thus obtained is an $`SO(3)`$-principal bundle over $`P`$ with $`3`$-Sasakian structure. Moreover, all three fibrations involved are Riemannian submersions. We recall now two basic moment maps of quaternion Kähler geometry. Let $`[h_0:h_1:\mathrm{}:h_n]`$ be the homogeneous coordinates on $`P^n`$: for each $`a=0,1,\mathrm{},n`$, we shall write the complex and real components of $`h_a`$ as follows: (2.2) $$h_a=z_a+w_aj=\alpha _a+\beta _ai+\gamma _aj+\delta _ak$$ where $`z_a=\alpha _a+\beta _ai,`$ and $`w_a=\gamma _a+\delta _ai.`$ The first moment map, induced by the diagonal action of $`U(1)`$ on $`P^n`$ is at the hyperkähler level of the bundle $`^{n+1}\backslash \{0\}P^n`$, given by: (2.3) $$\mu :^{n+1}\backslash \{0\},\mu =\underset{a=0}{\overset{n}{}}\overline{h}_aih_a.$$ On $`P^n`$, one has to regard the corresponding moment map $`\mu _{qK}`$ as a section of $`S^2H`$, the rank $`3`$ vector bundle of the compatible almost complex structures. In fact, by using the metric, $`\mu _{qK}`$ appears as a 2-form solution of $$d\mu _{qK}=i_X\mathrm{\Omega },$$ where $`\mathrm{\Omega }`$ is the Kähler 4-form and $`X`$ the Killing vector field generating the $`U(1)`$-action: , . However, since the zero level sets $`\mu ^1(0)^{n+1}`$ and $`\mu _{qK}^1(0)P^n`$ correspond to each other in the bundle projection $`^{n+1}\backslash \{0\}P^n`$, we shall refer to the definition $`\mu =_{a=0}^n\overline{h}_aih_a`$, whose zero set makes sense also when the $`h_a`$ are the homogeneous coordinates of $`P^n`$ . The reduced manifold $`\mu _{qK}^1(0)/U(1)`$ turns out to be the quaternion Kähler Wolf space $`\frac{SU(n+1)}{S(U(n1)\times U(2))}Gr_2(^{n+1})`$ , . The second moment map to be considered, generated by the action of $`Sp(1)`$, is (again at the hyperkähler level): (2.4) $$\nu :^{n+1}\backslash \{0\}^3,\nu =(\underset{a=0}{\overset{n}{}}\overline{h}_aih_a,\underset{a=0}{\overset{n}{}}\overline{h}_ajh_a,\underset{a=0}{\overset{n}{}}\overline{h}_akh_a),$$ and its corresponding quaternion Kähler moment map $`\nu _{qK}`$ can be viewed as a triple of 2-forms associated to the frame of Killing vector fields defining the $`Sp(1)`$-action. The corresponding reduced manifold $`\nu _{qK}^1(0)/Sp(1)`$ is now the Wolf space $`\frac{SO(n+1)}{SO(n3)\times SO(4)}\stackrel{~}{Gr}_4(^{n+1})`$, the Grassmannian of oriented four-planes of $`^{n+1}`$. ## 3. Statement of results ###### Theorem 3.1. (i) The focal set $`Foc_{P^n}P^n`$ coincides with $`\mu ^1(0)`$, and it is isometric to the total space of the induced Hopf $`S^1`$-bundle via the Plücker embedding $`Gr_2(^{n+1})P^N`$. (ii) The metric $`g_1`$, induced on $`\mu ^1(0)`$ by the Plücker embedding allows us to define on $`\mu ^1(0)`$ a Sasakian Einstein metric $`g`$. (iii) The Stiefel manifold $`V_2(^{n+1})`$ of orthonormal 2-frames in $`^{n+1}`$, diffeomorphic to the total space of the induced Hopf $`S^3`$-bundle via the embedding $`\mu ^1(0)P^n`$, admits an (integrable) complex structure $`J`$, not compatible with the standard hypercomplex structure of $`V_2(^{n+1})`$. We have a similar statement regarding the moment map $`\nu `$ induced by the action of $`Sp(1)`$ on $`P^n`$. We need to consider the following mutually congruent, totally geodesic embeddings of $`P^n`$ in $`P^n`$. (3.1) $$\begin{array}{cc}\hfill P_i^n=& \{hP^n;\gamma _a=\delta _a=0,a=1,\mathrm{},n\},\hfill \\ \hfill P_j^n=& \{hP^n;\beta _a=\delta _a=0,a=1,\mathrm{},n\},\hfill \\ \hfill P_k^n=& \{hP^n;\beta _a=\gamma _a=0,a=1,\mathrm{},n\}:\hfill \end{array}$$ here $`P_i^n`$ is the standard $`P^n`$ appearing in Theorem 3.1. ###### Theorem 3.2. (i) The zero level set $`\nu ^1(0)`$ coincides with the intersection $`M=Foc_{P^n}P_i^nFoc_{P^n}P_j^nFoc_{P^n}P_k^n`$ and is isometric to the total space of the induced Hopf $`S^1`$-bundle over the Fano manifold $`Z_{\stackrel{~}{Gr}_4(^{n+1})}`$, via the embeddings $`Z_{\stackrel{~}{Gr}_4(^{n+1})}Gr_2(^{n+1})P^N`$, the first of which is defined by regarding $`Z_{\stackrel{~}{Gr}_4(^{n+1})}`$ as the space of totally isotropic two-planes in $`^{n+1}`$. This isometry allows the construction of a Sasakian Einstein metric on $`\nu ^1(0)`$, and identifies it with the homogeneous 3-Sasakian manifold $`SO(n+1)/(SO(n3)\times Sp(1))`$. (ii) The Stiefel manifold $`\stackrel{~}{V}_4(^{n+1})`$ admits an (integrable) complex structure, projecting to the complex structure of $`Z_{\stackrel{~}{Gr}_4(^{n+1})}`$. Here $`Z_{\stackrel{~}{Gr}_4(^{n+1})}`$ is the twistor space of the quaternion Kähler Wolf space $`\stackrel{~}{Gr}_4(^{n+1})`$ given by the $`Sp(1)`$ reduction on $`P^n`$. $`Z_{\stackrel{~}{Gr}_4(^{n+1})}`$ is known to be a complex submanifold of the Kähler-Einstein Grassmannian $`Gr_2(^{n+1})`$, see for example p. 14 or p. 702. Statements 3.1, 3.2 give, in particular, some fibrations that can be collected into a diagram as follows. Here $`V_k(^{n+1})`$ and $`\stackrel{~}{V}_k(^{n+1})`$ denote the Stiefel manifolds of $`k`$-frames in $`^{n+1}`$ and of oriented $`k`$-frames in $`^{n+1}`$, respectively. ###### Corollary 3.1. There is a commutative diagram $$\begin{array}{cccccc}\stackrel{~}{V}_4(^{n+1})& & V_2(^{n+1})& & S^{4n+3}& \\ S^3& & S^3& & S^3& \\ \nu ^1(0)& & \mu ^1(0)& & P^n& S^{2N+1}\hfill \\ S^1& & S^1& & & \hfill \\ Z_{\stackrel{~}{Gr}_4(^{n+1})}& & Gr_2(^{n+1})& & P^N& \\ S^2& & & & & \\ \stackrel{~}{Gr}_4(^{n+1})& & & & & \end{array}$$ of principal $`S^1`$ and $`S^3`$-bundles and Riemannian submersions. The zero level sets $`\mu ^1(0)`$ and $`\nu ^1(0)`$ are thus total spaces of induced Hopf $`S^1`$-bundles and, as such, are minimal submanifolds of the sphere $`S^{2N+1}`$. ## 4. The proofs ### Proof of Theorem 3.1 (i) and (ii) The first observation is: $`Foc_{P^1}P^1`$ is the level set of the moment map associated to the standard $`U(1)`$-action on $`P^1`$. In the homogeneous coordinates $`[h_0:h_1]`$ of $`P^1`$, using the notations from formula (2.1), $`P^1`$ is given by: (4.1) $$w_0=0,w_1=0\text{or}\gamma _0=\delta _0=\gamma _1=\delta _1=0.$$ The pair $`P^1P^1`$ can be identified with $`S^2S^4`$, and the focal set of a totally geodesic sphere $`S^p`$ in $`S^n`$ is the unit $`S^{np1}`$ in the $`(np)`$-dimensional orthogonal complement of the $`^{p+1}`$ containing $`S^p`$ (, p. 286). This identification can be made explicit by using the coordinate $`h=\alpha +\beta i+\gamma j+\delta k=h_0h_1^1`$ in the affine line $`h_10`$, whose corresponding real coordinates are: $$\begin{array}{cc}\hfill \alpha & =\frac{\alpha _0\alpha _1+\beta _0\beta _1+\gamma _0\gamma _1+\delta _0\delta _1}{r^2},\beta =\frac{\alpha _1\beta _0\alpha _0\beta _1\gamma _0\delta _1+\gamma _1\delta _0}{r^2},\hfill \\ \hfill \gamma & =\frac{\alpha _1\gamma _0\alpha _0\gamma _1\delta _0\beta _1+\beta _0\delta _1}{r^2},\delta =\frac{\alpha _1\delta _0\alpha _0\delta _1\gamma _1\beta _0+\gamma _0\delta _1}{r^2},\hfill \end{array}$$ where $`r^2=\alpha _1^2+\beta _1^2+\gamma _1^2+\delta _1^2`$. If $`(x_1,\mathrm{},x_5)`$ are the standard coordinates on $`^5`$, the (inverse) stereographic projection $`^4S^4`$ reads: (4.2) $$(\alpha ,\beta ,\gamma ,\delta )(x_1,\mathrm{},x_5)=(\frac{2\alpha }{r^2+1},\frac{2\beta }{r^2+1},\frac{2\gamma }{r^2+1},\frac{2\delta }{r^2+1},\frac{r^21}{r^2+1}).$$ Thus, if $`x_3=x_4=0`$ are the equations of a totally geodesic $`S^2`$ in $`S^4`$, the focal set of $`S^2`$ in $`S^4`$ is the $`S^1`$ given by the equations $`x_1=x_2=x_5=0`$, corresponding to (4.3) $$\alpha =\beta =0,\gamma ^2+\delta ^2=1.$$ On the other hand, the moment map $`\mu `$ on $`P^1`$ can be written in the non-homogeneous coordinate $`h=h_0h_1^1`$ as $`\mu =\overline{h}ih+i`$. Its level set $`\mu ^1(0)`$ is thus: (4.4) $$\alpha _1^2+\beta _1^2\gamma _1^2\delta _1^2+1=0,\alpha \delta \beta \gamma =0,\alpha \gamma +\beta \gamma =0,$$ and the systems (4.3) and (4.4) are equivalent. We now treat the case $`n>1`$. From the definition of the focal set, we only need to look at geodesics normal to $`P^n`$. These are normal to all the complex projective lines $`L^1P^n`$, thus their focal points with respect to $`P^n`$ are also focal points for all the lines $`L^1`$ contained in $`P^n`$. Any such line $`L^1`$ in $`P^n`$ belongs to a unique quaternionic projective line $`L^1`$ and the latter is totally geodesic in $`P^n`$. It follows that any geodesic that is normal to a $`L^1`$ and tangent to the corresponding $`L^1`$ at a given point, remains tangent to $`L^1`$ for its entire length. Hence, if we show that through any point $`x`$ of $`P^n`$ and for any $`vT_x^{}P^n`$ there exists a projective quaternionic line $`L^1`$ containing $`v`$, we may deduce: $`Foc_{P^n}P^n`$ is the union of all the focal sets $`Foc_{L^1}L^1`$. This can be seen from the diagram $$\begin{array}{ccc}^{n+1}\backslash \{0\}& & ^{n+1}\backslash \{0\}\\ & & \\ P^n& & P^n.\end{array}$$ by looking at the complex planes $`L_2^{}`$ in $`^{n+1}`$ containing the fibre $`^{}`$ and the corresponding hypercomplex 2-planes $`L_2^{}`$ in $`^{n+1}`$ containing the fibre $`^{}`$. Observe that, for any vector $`\stackrel{}{v}^{n+1}`$ which is normal to the standard embedded $`^{n+1}`$, there exist a $`L_2^{}`$ and a $`L_2^{}`$ with $`\text{span}_{}\{L_2^{},\stackrel{}{v}\}L_2^{}`$: indeed, if $`^{n+1}=\text{span}_{}\{\stackrel{}{e_0},\mathrm{},\stackrel{}{e_n}\}`$ and $`^{n+1}=\text{span}_{}\{\stackrel{}{e_0},\mathrm{},\stackrel{}{e_n}\}`$, we have: $$\stackrel{}{v}^{n+1}\text{if and only if}\stackrel{}{v}=\underset{a=0}{\overset{n}{}}\lambda _aj\stackrel{}{e_a}+\mu _ak\stackrel{}{e_a}.$$ Hence, if $$\stackrel{}{w}=j\stackrel{}{v}=\underset{a=0}{\overset{n}{}}\lambda _a\stackrel{}{e_a}\mu _ai\stackrel{}{e_a},$$ we obtain $$L_2^{}=\text{span}_{}\{\stackrel{}{e_0},\stackrel{}{w}\}$$ and $$L_2^{}=\text{span}_{}\{\stackrel{}{e_0},i\stackrel{}{e_0},j\stackrel{}{e_0},k\stackrel{}{e_0},\stackrel{}{w},i\stackrel{}{w},j\stackrel{}{w},k\stackrel{}{w}\}$$ satisfy the condition. The identification of $`Foc_{P^n}P^n`$ with $`\mu ^1(0)`$ is then completed by the following observation: The subset $`\mu ^1(0)P^n`$ is the union of the zero sets of the moment maps $`\mu _1`$ associated to the standard $`U(1)`$-action on all the projective quaternionic lines $`L^1P^n`$. Let $`\{p_0,p_1,\mathrm{},p_n\}`$ be the canonical frame of $`P^n`$ with unit point $`u`$ and let $`[h_0:\mathrm{}:h_n]`$ be the homogeneous coordinates with respect to this frame. Accordingly, the moment map reads $`\mu (h)=_{a=0}^n\overline{h}_aih_a`$. Fix a $`q\mu ^1(0)`$, *i.e.* $`_{a=0}^n\overline{q}_aiq_a=0`$, and note that $`q`$ cannot be real; however, we may suppose $`q_00`$. Let $`L^1P^1`$ be the quaternionic projective line through $`q`$ and $`p_1`$. To compare the intersection $`\mu ^1(0)L^1`$ with the zero level set of the moment map in $`L^1P^1`$, we change the frame in $`P^n`$ as follows. We want new homogeneous coordinates $`[k_0:k_1:\mathrm{}:k_n]`$ such that $`q=[1:j:0:\mathrm{}:0]`$ and the coordinates of $`p_1`$,…, $`p_n`$ remain unchanged. With the new unit point $`v=[q_0:1:1:1:\mathrm{}:1]`$, the matrix of the change of coordinates is: $$A=\left(\begin{array}{ccccc}q_0& 0& 0& \mathrm{}& 0\\ q_1j& 1& 0& \mathrm{}& 0\\ q_2& 0& 1& \mathrm{}& 0\\ 5\\ q_n& 0& 0& \mathrm{}& 1\end{array}\right).$$ Thus, $`{}_{}{}^{t}[h_0:\mathrm{}:h_n]=A^t[k_0:\mathrm{}:k_n]`$, and the moment map is: $$\mu =\underset{a=0}{\overset{n}{}}\overline{h_a}ih_a=\overline{k_0}(\underset{a=0}{\overset{n}{}}\overline{q_a}iq_a)k_0\overline{k_0}jijk_0\underset{b=1}{\overset{n}{}}\overline{k_b}ik_b=\underset{b=0}{\overset{n}{}}\overline{k_b}ik_b.$$ It follows that $`\mu ^1(0)`$ is invariant under these projective changes of coordinates. As the coordinates on $`L^1`$ are $`[k_0:k_1]`$, its moment map is $`\mu _1=\overline{k_0}ik_0\overline{k_1}ik_1`$ with the same level set $`\mu _1^1(0)`$ described for $`P^1`$. This establishes the inclusion $`\mu ^1(0)|_{L^1}\mu _1^1(0)`$. As the converse inclusion is clear, for any projective line $`L^1`$ in $`P^n`$, the proof of the identification $`Foc_{P^n}P^n=\mu ^1(0)`$ is complete. Next, we prove that $`Foc_{P^n}P^n`$ is isometric to the total space of the induced Hopf bundle over the Grassmannian $`Gr_2(^{n+1})`$. In fact, since $`Foc_{P^n}P^n`$ is simply connected (, p. 17), the existence of a diffeomorphism with the induced Hopf bundle is a consequence of the following observation: Let $`\pi :PGr_2(^{n+1})`$ be a principal circle bundle with simply connected $`P`$. Then $`P`$ is diffeomorphic to the total space of the induced Hopf bundle of $`S^{2N+1}P^N`$, $`N=\left(\genfrac{}{}{0pt}{}{n+1}{2}\right)1`$, via the Plücker embedding $`Gr_2(^{n+1})P^N`$. In fact, the principal $`S^1`$ -bundles over the base $`B`$ are classified by their Chern class in $`H^2(B,)`$. Since $`H^2(Gr_2(^{n+1}),)`$ is generated by the class of the Kähler form, one can denote by $`P_a`$ the $`S^1`$-bundle over $`Gr_2(^{n+1})`$ associated to $`a`$. Observe that $`P_1`$ (resp. $`P_1`$) is the circle bundle associated to the canonical line bundle $`O_{Gr_2(^{n+1})}(1)`$ (resp. its dual $`O_{Gr_2(^{n+1})}(1)`$). But dual complex line bundles are diffeomorphic as real vector bundles. Hence $`P_1`$ is diffeomorphic to $`P_1`$. Let us now show that if $`P_a`$ is simply connected, then $`a=\pm 1`$. From $`\pi _1(P)=0`$, we have $`H^1(P,)=0`$ and $`H^2(P,)`$ torsion free. Thus the Gysin sequence of the $`S^1`$-bundle $`\pi `$: $$\begin{array}{cc}\hfill 0H^1(P,)& H^0(Gr_2(^{n+1}),)H^2(Gr_2(^{n+1}),)\hfill \\ & H^2(P,)0\hfill \end{array}$$ reduces to: $$0\stackrel{c_1}{}H^2(P,)0,$$ where $`c_1`$ is the Chern class of $`\pi `$. Hence, $`c_1\pm 1`$ implies $`H^2(P,)_n`$ for some $`n2`$. Consequently, $`c_1`$ is, up to sign, the Chern class of the induced Hopf bundle. We now look at the metric $`g_1`$ inherited from $`(S^{2N+1},can)`$ by the total space $`Foc_{P^n}(P^n)`$ of the induced Hopf bundle and at the metric $`g`$ induced on $`Foc_{P^n}(P^n)`$ from $`P^n`$. Note that both $`(Foc_{P^n}(P^n),g_1)`$ and $`(Foc_{P^n}(P^n),g)`$ are Riemannian submersions with geodesic fibres $`S^1`$ over $`Gr_2(^{n+1})`$ (cf. for $`g_1`$ and for $`g`$). Thus Theorem 9.59 in , p. 249, can be used to conclude that $`g=g_1`$. This completes the proof of (i) and (ii) in Theorem 3.1. ###### Remark 4.1. An alternative way of recognizing that $`Foc_{P^n}P^n=\mu ^1(0)`$ is to look at the standard isometric action of $`SU(n+1)`$ on $`P^n`$ and to see that the two subsets are obtained as homogeneous spaces of $`SU(n+1)`$ with the same isotropy groups. The homogeneous space is in fact $`\frac{SU(n+1)}{SU(2)\times SU(n1)}`$, which was shown in , p. 17, to be a singular orbit of the action of $`SU(n+1)`$ (it was denoted there by $`Q^n`$). A similar description of $`\mu ^1(0)`$ as a homogeneous space is given in , p. 65, in relation with the problem of studying local compatible complex structures in $`P^n`$. Compare also with , p. 171, where $`\frac{SU(n+1)}{SU(2)\times SU(n1)}`$ is called the ”Grassmannian of *oriented* two-planes” of $`^{n+1}`$. ###### Remark 4.2. We proved in that the metric $`g_1`$ is Sasakian and $`\eta `$-Einstein, and this is in accordance with formulas following Proposition 9 in . As for the extrinsic geometry of $`Foc_{P^n}(P^n)`$, J. Berndt proves (Corollary 1 in ) that the immersion in $`P^n`$ is minimal. It is interesting to observe that the same holds for the immersion $`(Foc_{P^n}P^n,g_1)`$ in $`(S^{2N+1},can)`$, see . ### Proof of Theorem 3.2 (i) The identification of $`\nu ^1(0)`$ with the intersection $`M`$ is an immediate consequence of the definition of $`\nu `$ and of the first statement of Theorem 3.1. To see that $`\nu ^1(0)`$ is isometric to the induced Hopf bundle discussed above, note that the last observation in the proof of Theorem 3.1 (i) and (ii) still holds for principal circle bundles with simply connected $`P`$ over any complex algebraic projective submanifold $`B`$ of $`P^N`$ with $`H^2(B,)`$. This applies in particular to $`B=Z_{\stackrel{~}{Gr}_4(^{n+1})}`$, as soon as one recognizes that $`\nu ^1(0)`$ is simply connected. To see this, observe that $`\nu ^1(0)`$ can be regarded as the homogeneous space $`\frac{SO(n+1)}{SO(n3)\times Sp(1)}`$ via the transitive action of $`SO(n+1)`$ on $`\nu ^1(0)`$. This last action comes, in fact, from the natural action of $`SO(n+1)`$ on $`P^n`$ and J. Berndt’s observation that $`SU(n+1)`$ acts transitively on $`Foc_{P^n}(P^n)`$. His argument can be triplicated to produce transitive actions of groups $`SU_i(n+1)`$, $`SU_j(n+1)`$, $`SU_k(n+1)`$ (the first one is the standard $`SU(n+1)`$ in $`Sp(n+1)`$, the other two are similarly defined by interchanging the rôles of the unit quaternions $`i`$, $`j`$, $`k`$) on the zero level sets $`\mu _i^1(0)`$, $`\mu _j^1(0)`$, $`\mu _k^1(0)`$. The isotropy subgroup of the action on $`\nu ^1(0)`$ at the point $`[1:i:j:k:0\mathrm{}:0]`$ is then $`Sp(1)\times SO(n3)`$. The homotopy sequence associated to the homogeneous manifold $`\frac{SO(n+1)}{Sp(1)\times SO(n3)}`$ thus shows that $`\nu ^1(0)`$ is simply connected, so that it is diffeomorphic with the total space of the induced Hopf bundle, now over the Fano manifold $`Z_{\stackrel{~}{Gr}_4(^{n+1})}`$. As an induced $`S^1`$-Hopf bundle, $`\nu ^1(0)`$ has a Sasakian $`\eta `$ \- Einstein structure. This can be deformed to a Sasakian-Einstein metric, which is still a Riemannian submersion after rescaling the standard Kähler-Einstein metric of $`Z_{\stackrel{~}{Gr}_4(^{n+1})}`$ . On the other hand, the composition of the fiberings $$\nu ^1(0)\stackrel{S^1}{}Z_{\stackrel{~}{Gr}_4(^{n+1})}\stackrel{S^2}{}\stackrel{~}{Gr}_4(^{n+1})$$ is an $`SO(3)`$-bundle which endows $`\nu ^1(0)`$ with a $`3`$-Sasakian structure via the inversion Theorem 4.6 of . ### Proof of Theorem 3.1 (iii) and Theorem 3.2 (ii) In both cases we construct a complex structure on the total space of an $`S^3`$ bundle over a Sasakian manifold, induced by the Hopf bundle $`S^{4n+3}P^n`$. Both Stiefel manifolds under discussion are induced Hopf $`S^3`$-bundles, as is easily recognized by regarding them as homogeneous spaces. Thus the induced homogeneous metrics make them Riemannian submersions with fibers $`S^3`$. More generally: ###### Proposition 4.1. Let $`\pi :PB`$ be a principal $`S^3`$-bundle induced by the Hopf bundle $`S^{4n+3}P^n`$, and let $`g^B`$, $`^B`$ be the induced metric and Levi Civita connection on $`BP^n`$. Assume that $`B`$ admits a Killing vector field $`\xi `$ such that $`\phi =^B\xi `$ defines on $`B`$ a Sasakian structure Then $`P`$ admits an almost complex structure. ###### Proof. Let $`g^P`$ be the induced metric on $`PS^{4n+3}`$, so that $`\pi `$ is a Riemannian submersion. For any $`X𝒳(B)`$ we let $`X^{}`$ be its horizontal lift on $`P`$. Let $`\xi _1`$, $`\xi _2`$, $`\xi _3`$ be the unit Killing vector fields which give the usual $`3`$-Sasakian structure of the fibers $`S^3`$ and $`\eta _1`$, $`\eta _2`$, $`\eta _3`$ their duals with respect to the canonical metric of $`S^3`$. The $`\xi _i`$ may be viewed as vector fields on $`P`$. Let $`\widehat{\eta }_i`$ be their dual forms with respect to the metric $`g^P`$. Restricted to any fibre, the $`\widehat{\eta }_i`$ coincide with the $`\eta _i`$. The usual splitting $`TP=𝒱`$ into vertical and horizontal parts may be refined to: $$TP=\text{span}\{\xi _1,\xi _2,\xi _3\}\text{span}\{\xi ^{}\}^{},$$ where $`^{}`$ represents the horizontal vector fields orthogonal to the horizontal lift $`\xi ^{}`$ of $`\xi `$.. Define the almost complex structure $`J`$ on $`P`$ by: * $`J\xi _1=\xi _2,J\xi _2=\xi _1`$, * $`J\xi _3=\xi ^{},J\xi ^{}=\xi _3`$, * $`JX^{}=(\phi X)^{}`$ for any $`X𝒳(B)`$ orthogonal to $`\xi `$. Note that for $`X\xi `$, $`X^{}`$ is a section of $`^{}`$. As the restriction of $`\phi `$ to $`\xi ^{}`$ is an endomorphism of $`\xi ^{}`$, the last item in the definition is consistent. One easily shows that $`J^2=1`$ and $`J`$ is compatible with $`g`$. ∎ To discuss the integrability of the constructed $`J`$, we follow the discussion developed for an almost hypercomplex structure, computing the Nijenhuis tensor field on all possible combinations of vertical and/or horizontal vector fields. Note first that the horizontal distribution $``$ is an $`sp(1)`$-connection in the induced Hopf $`S^3`$ bundle $`PB`$. This follows from the fact that the bracket of any horizontal $`X^{}`$ with a vertical vector field is horizontal, a consequence of the Killing property of the $`\xi _i`$ with respect to $`g^P`$. Thus, in particular we get that for $`i=1,2,3`$: (4.5) $$\widehat{\eta }_k[\xi _i,X^{}]=\widehat{\eta }_k[\xi _i,\xi ^{}]=0,$$ for $`k=1,2,3`$. Now we can prove: ###### Proposition 4.2. Assume that the curvature form $`\mathrm{\Omega }`$ of the $`sp(1)`$-connection $``$ satisfies the following conditions: 1) $`\mathrm{\Omega }((\phi X)^{},(\phi Y)^{})=\mathrm{\Omega }(X^{},Y^{})`$, *i.e.* $`\mathrm{\Omega }`$ is of type $`(1,1)`$ with respect to $`J`$, 2) $`\mathrm{\Omega }(X^{},\xi ^{})=0`$ for any $`X\xi `$. Then the almost complex structure $`J`$ is integrable. ###### Remark 4.3. Conditions 1), 2) in the former Proposition express a compatibility between the Sasakian structure of the base and the bundle structure of $`P`$. Since the vertical components of $`\mathrm{\Omega }`$ are precisely the $`d\widehat{\eta }_i`$, the two conditions give corresponding equations for $`d\widehat{\eta }_i`$. Moreover, the condition $`d\widehat{\eta }_i((\phi X)^{},(\phi Y)^{})=d\widehat{\eta }_i(X^{},Y^{})`$ is easily checked to be equivalent with $`d\widehat{\eta }_i((\phi X)^{},Y^{})+d\widehat{\eta }_i(X^{},(\phi Y)^{})=0`$. Now we can give the proof of the Proposition 4.2: ###### Proof. We compute the Nijenhuis tensor field of $`J`$: $$[J,J](A_1,A_2)=[A_1,A_2]+J[JA_1,A_2]+J[A_1,JA_2][JA_1,JA_2]$$ for all possible pairs $`(A_1,A_2)`$, noting that, due to the tensorial character of $`[J,J]`$, when dealing with horizontal (resp. vertical) vector fields it is enough to work with basic ones (resp. with $`\xi _i`$, $`i=1,2,3`$). 1. Let first $`A_1=X^{}`$, $`A_2=Y^{}`$, $`X,Y\xi `$. Denoting by $`\widehat{\eta }`$ the dual of $`\xi ^{}`$ we get: (4.6) $$[X^{},Y^{}]=[X^{},Y^{}]^{}+\widehat{\eta }([X^{},Y^{}])\xi ^{}+\text{vertical part}.$$ where the denotes the $`^{}`$ part. By $`\pi `$-corelation, $`[X^{},Y^{}]=[X,Y]^{}^{}`$. Moreover, from $`\widehat{\eta }(X^{})=\widehat{\eta }(Y^{})=0`$ we get $`\widehat{\eta }([X^{},Y^{}])=d\widehat{\eta }(X^{},Y^{}).`$ The vertical part of $`[X^{},Y^{}]`$ must be of the form $`_{i=1}^3\widehat{\eta }_i([X^{},Y^{}])\xi _i`$. Hence: (4.7) $$[X^{},Y^{}]=[X^{},Y^{}]^{}d\widehat{\eta }(X^{},Y^{})\xi ^{}\underset{i=1}{\overset{3}{}}d\widehat{\eta }(X^{},Y^{})\xi _i.$$ By similar computations: $$\begin{array}{cc}\hfill J[JX^{},Y^{}]& =(\phi [\phi X,Y])^{}{}_{}{}^{}+d\widehat{\eta }((\phi X)^{},Y^{})\xi _3\hfill \\ & d\widehat{\eta }_1((\phi X)^{},Y^{})\xi _2+d\widehat{\eta }_2((\phi X)^{},Y^{})\xi _1d\widehat{\eta }_3((\phi X)^{},Y^{})\xi ^{},\hfill \\ \hfill J[X^{},JY^{}]& =(\phi [X,\phi Y])^{}{}_{}{}^{}+d\widehat{\eta }(X^{},(\phi Y)^{})\xi _3\hfill \\ & d\widehat{\eta }_1(X^{},(\phi Y)^{})\xi _2+d\widehat{\eta }_2(X^{},(\phi Y)^{})\xi _1d\widehat{\eta }_3(X^{},(\phi Y)^{})\xi ^{},\hfill \\ \hfill [JX^{},JY^{}]& =[\phi X,\phi Y]^{}{}_{}{}^{}d\widehat{\eta }((\phi X)^{},(\phi Y)^{})\xi ^{}\hfill \\ & \underset{i=1}{\overset{3}{}}d\widehat{\eta }_i((\phi X)^{},(\phi Y)^{})\xi _i.\hfill \end{array}$$ Hence, using the $`(1,1)`$ character of $`d\widehat{\eta }_i`$ and Remark 4.3, we find $$[J,J](X^{},Y^{})=[\phi X,\phi Y]^{}{}_{}{}^{}\{d\widehat{\eta }(X^{},Y^{})d\widehat{\eta }((\phi X)^{},(\phi Y)^{})\}\xi ^{}.$$ As we know $`[\phi X,\phi Y]+2d\eta (X,Y)\xi =0`$ (this is the normality condition of the Sasakian structure of $`B`$) the horizontal lift of this (null) tensor field is zero, hence also its component in $`^{}`$ is zero. But this is precisely $`[\phi X,\phi Y]^{}^{}`$. On the other hand, on any Sasakian manifold one has: $$d\eta (X,Y)=g(X,\phi Y),\phi ^2X=X+\eta (X)\xi ,$$ thus $`d\eta (X,\phi Y)+d\eta (\phi X,Y)=0`$ and $`d\eta (\phi X,\phi Y)d\eta (X,Y)=0`$. By horizontally lifting these equations we obtain the annulation of the $`\xi ^{}`$ component, hence $`[J,J](X^{},Y^{})=0`$. 2. Let now $`A_1=X^{}`$, $`A_2=\xi ^{}`$ ($`X\xi `$). Then: $$\begin{array}{cc}\hfill [J,J](X^{},\xi ^{})& =[X^{},\xi ^{}]+J[JX^{},\xi ^{}]+J[X^{},J\xi ^{}][JX^{},J\xi ^{}]=\hfill \\ & =[X^{},\xi ^{}]+J[(\phi X)^{},\xi ^{}]J[X^{},\xi _3]+[(\phi X)^{},\xi _3]=\hfill \\ & =[X^{},\xi ^{}]+J[(\phi X)^{},\xi ^{}],\hfill \end{array}$$ as the last two brackets are zero by (4.5). As above, using Remark 4.3 and $`d\widehat{\eta }_i(X^{},\xi ^{})=0`$ (condition 2) in the statement) we obtain: $$[J,J](X^{},\xi ^{})=([X,\xi ]+\phi [\phi X,\xi ])^{}{}_{}{}^{}=0,$$ because, as on a Sasakian manifold $`\phi \xi =0`$, we can add in the paranthesis the terms $`[X,\phi \xi ][\phi X,\phi \xi ]`$ obtaining $`([X,\xi ]+\phi [\phi X,\xi ]+\phi [X,\phi \xi ][\phi X,\phi \xi ])^{}^{}`$ = $`([\phi ,\phi ](X,\xi ))^{}{}_{}{}^{}=0`$ by the normality condition on $`B`$. 3. We now choose $`A_1=X^{}`$ and $`A_2=\xi _i`$ ($`i=1,2`$). We have: $$[J,J](X^{},\xi _1)=[X^{},\xi _1]+J[JX^{},\xi _1]+J[X^{},J\xi _1][JX^{},J\xi _1]=0,$$ because $``$ is a $`sp(1)`$-connection. Similarly for $`[J,J](X^{},\xi _2)=0.`$ 4. For $`A_1=X^{}`$ and $`A_2=\xi _3`$ we find: $$\begin{array}{cc}\hfill [J,J](X^{},\xi _3)& =[X^{},\xi _3]+J[JX^{},\xi _3]+J[X^{},J\xi _3][JX^{},J\xi _3]=\hfill \\ & =J[X^{},\xi ^{}][(\phi X)^{},\xi ^{}],\hfill \end{array}$$ the brackets with $`\xi _3`$ being zero by (4.5). The horizontal component of the remaining two brackets is $`([\phi [X,\xi ][\phi X,\xi ])^{}{}_{}{}^{}d\widehat{\eta }_3(X^{},\xi ^{})\xi _i`$. The $`\xi ^{}`$ component as well as the vertical component vanish by assumption 2) in the statement. Finally, using Sasakian identities and the normality condition on $`B`$ we have $`([\phi [X,\xi ][\phi X,\xi ])^{}{}_{}{}^{}=0`$. 5. Immediate computation shows that in the remaining ”mixed” case $`[J,J](\xi _i,\xi ^{})=0`$. 6. We are left with the computation of $`[J,J]`$ on vertical fields. Obviously $`[J,J](\xi _1,\xi _2)=0`$. Then $$\begin{array}{cc}\hfill [J,J](\xi _1,\xi _3)& =[\xi _1,\xi _3]+J[J\xi _1,\xi _3]+J[\xi _1,J\xi _3][J\xi _1,J\xi _3]\hfill \\ & =[\xi _1,\xi 3]+J[\xi _2,\xi _3]+J[\xi _1,\xi ^{}][\xi _2,\xi ^{}]=0\hfill \end{array}$$ by $`[\xi ,\xi _j]=2\epsilon _{ijk}\xi _k`$ and (4.5). Same arguments show that $`[J,J](\xi _2,\xi _3)=0`$, thus completing the proof. ∎ ###### Remark 4.4. The Kähler form of $`(P,g,J)`$ is $$\omega =d\pi ^{}\eta +\pi ^{}\eta \eta _3d\eta _3;$$ this shows that $`d\omega 0`$, hence the structure is not Kählerian. A similar computation proves that $`L_\xi ^{}J=L_{\xi _3}J=0`$, thus $`\xi ^{}`$ and $`\xi _3`$ are infinitesimal automorphisms of the constructed complex structure. ###### Remark 4.5. The complex structure $`J`$ on $`P`$ depends on the choice of a $`3`$-Sasakian structure on the fibre. But one can see that different choices of $`3`$-Sasakian triples $`\{\xi _1,\xi _2,\xi _3\}`$ produce complex structures that are conjugated in $`End(TP)`$. We can now go back to the Stiefel manifolds $`V_2(^{n+1})`$ and $`\stackrel{~}{V}_4(^{n+1})`$, and complete the proof of Theorems 3.1 (iii) and 3.2 (ii). We have just to verify that the induced Hopf bundles $`V_2(^{n+1})\mu ^1(0)`$ and $`\stackrel{~}{V}_4(^{n+1})\nu ^1(0)`$ inherit from the inclusions $`\nu ^1(0)\mu ^1(0)P^n`$ horizontal distributions $``$ satisfying the curvature properties of Proposition 4.2. Now property 1) simply express that the $`sp(1)`$-connection given by $``$ is part of a $`u(2)`$-connection in the bundles $`V_2(^{n+1})Gr_2(^{n+1})`$ and $`\stackrel{~}{V}_4(^{n+1})Z_{\stackrel{~}{G}r_4(^{n+1})}`$, a fact easily recognized as in the case of hypercomplex structures in $`V_2(^{n+1})`$ (cf., proof of Thm 1.10, as well as , Thm. 1.13). The meaning of property 2) is that the curvature of such a $`sp(1)`$-connection is given by a (1,1)-form with respect to the almost complex structure $`J`$. This follows for example from the computation carried out in , p. 63, for the Hopf bundle. One has to take into account that the rôle of the $`U(1)`$ and of the $`Sp(1)`$-actions on both Stiefel manifolds correspond to the standard basic choice of left and right multiplication by scalars on $`^{n+1}`$. Then both on $`V_2(^{n+1})`$ and on $`\stackrel{~}{V}_4(^{n+1})`$ the almost complex structure $`J`$ satisfies the compatibility conditions with the Sasakian structures of $`\mu ^1(0)`$ and $`\nu ^1(0)`$, as expressed by conditions 1) and 2) of Proposition 4.2. Note that the complex structure obtained in this way on $`V_2(^{n+1})`$ projects to the complex structure underlying the Kähler structure of $`Gr_2(^{n+1})`$. This latter is well known to be not compatible with the quaternion Kähler structure of this Grassmannian. But it is precisely this quaternion Kähler structure which is lifted to a $`3`$-Sasakian structure and then, by means of an appropriate circle bundle, to a hypercomplex structure on $`V_2(^{n+1})`$, cf. , . Hence our complex structure is not compatible with the standard hypercomplex one of $`V_2(^{n+1})`$. This completes the proof of Theorem 3.1 (iii) and 3.2 (ii). ### Proof of Corollary 3.1 It remains only to show that the total space of an induced $`S^1`$ Hopf bundle is minimal in $`(S^{2N1},can)`$. In fact, more generally, in a commutative diagram $$\begin{array}{ccc}\overline{N}& \stackrel{\overline{i}}{}& \overline{M}\\ \pi _N& & \pi _M& & \\ N& \stackrel{i}{}& M\end{array}$$ of immersions $`i`$, $`\overline{i}`$ and Riemannian submersions $`\pi _N`$, $`\pi _M`$ with totally geodesic fibres, we see that $`N`$ is minimal in $`M`$ if and only if $`\overline{N}`$ is minimal in $`\overline{M}`$. This follows by a direct computation of the mean curvature vector fields of $`i`$ and $`\overline{i}`$ using the Gauss formula of a submanifold and formula (9.25 a) in . ## 5. Further Observations As mentioned in the introduction, the zero level set $`\mu ^1(0)P^2`$ is diffeomorphic to a sphere $`S^5`$ and the projection to the reduced manifold can be identified with the Hopf fibration $`S^5P^2`$. Going to the next case $`n=3`$, the projection $`\mu ^1(0)Gr_2(^4)`$, now an induced Hopf fibration, can be described by looking at the families of submanifolds in $`Gr_2(^4)`$ that are either Kähler-Einstein or quaternion Kähler. The Grassmannian $`Gr_2(^4)`$ admits in fact a natural complex Kähler structure, as well as two distinct quaternion Kähler structures induced via the isomorphism of vector bundles $`TGr_2(^4)VV^{}`$ from the (almost) hypercomplex structure on the tautological vector bundle $`V`$ or on its orthogonal complement $`V^{}`$ (cf. ). The families of submanifolds we want to look at on $`Gr_2(^4)`$ are described as follows (see , , , ). There are two families $``$, $`^{}`$ of complex projective planes $`𝐂P^2`$, $`𝐂P^2^{}`$, a family $`^{\prime \prime }`$ of products $`P^1\times P^1`$, and a family $`^{\prime \prime \prime }`$ of spheres $`S^4`$; they are given by: $$\begin{array}{cc}\hfill =& \{\text{planes contained in a }3\text{-space of }^4\},\hfill \\ \hfill ^{}=& \{\text{planes through a line of }^4\},\hfill \\ \hfill ^{\prime \prime }=& \{\text{planes that are invariant for a hypercomplex}\hfill \\ & \text{structure }J\text{ of }^4\},\hfill \\ \hfill ^{\prime \prime \prime }=& \{\text{planes given by pairs of lines in two fixed orthogonal}\hfill \\ & \text{planes of }^4\}.\hfill \end{array}$$ All these $`4`$-dimensional submanifolds of $`Gr_2(^4)`$ have nice intersection properties, some of which can also be formulated in terms of projective geometry of the lines in the 3-dimensional space $`P^3`$, the context where the Klein quadric $`Q^4P^5`$, isometric to $`Gr_2(^4)`$, was first introduced. Instead of listing these intersection properties on $`Gr_2(^4)`$ (cf. , pp. 508-512, for some of them; the remaining ones can be similarly deduced), we formulate the corresponding properties for the families of 5-dimensional submanifolds obtained at the Sasakian Einstein level as induced Hopf bundles over the members of families $``$,…, $`^{\prime \prime \prime }`$. ###### Proposition 5.1. The $`9`$-dimensional Sasakian Einstein focal set $`Foc_{P^3}P^3\mu ^1(0)P^3`$ contains the following families of 5-dimensional submanifolds, each of which fibers in circles over a complex or a quaternionic submanifold of $`Gr_2(^4)`$. There are two families $``$, $`^{}`$ of Sasakian 5-spheres $`S^5`$, $`S_{}^{}{}_{}{}^{5}`$, a family $`^{\prime \prime }`$ of Sasakian products $`S^3\times S^2`$ and a family $`^{\prime \prime \prime }`$ of products $`S^4\times S^1`$. The induced metrics on members of the families $``$, $`^{}`$,$`^{\prime \prime }`$ are Sasakian $`\eta `$-Einstein, and can be modified to Sasakian Einstein metrics by formula (2.1). The intersection properties of these 5-dimentional submanifolds are the following. (a) Pairs of 5-spheres in the same family intersect in a circle, and pairs of 5-spheres in different families either do not intersect or intersect in an $`S^3`$. (b) Pairs of submanifolds of type $`S^3\times S^2`$ intersect either in an $`S^3`$ or in a pair of disjoint circles. A submanifold of type $`S^3\times S^2`$ intersects a 5-sphere in an $`S^3`$. (c) Any submanifold of type $`S^4\times S^1`$ intersects a $`5`$-sphere in a circle, intersects an $`S^3\times S^2`$ in two disjoint circles, and any pair of submanifolds of type $`S^4\times S^1`$ intersect in two disjoint circles. This kind of geometry of submanifolds, now described for the level set $`\mu ^1(0)P^3`$, can be formulated for all the zero level sets of moment maps appearing in the diagram of Corollary 3.1. There is such a level set for each odd dimension. For example, $`\nu ^1(0)P^4`$ is diffeomorphic to a sphere $`S^7`$, yielding as reduced manifold $`P^1`$ and fibering in circles over its twistor space $`P^3`$, a Kähler submanifold of the Grassmannian $`Gr_2(^5)`$. Thus Sasakian ($`\eta `$)-Einstein submanifolds of type $`S^5`$ and $`S^3\times S^2`$ can be determined in $`S^7`$ (cf. ), and intersection properties like in Proposition 5.1 are obtained. The 11-dimensional example is the level set $`\nu ^1(0)P^5`$, identified in Theorem 3.2 with an intersection of three focal sets in $`P^5`$. This 11-dimensional manifold is diffeomorphic to the (unique) 3-Sasakian homogeneous manifold projecting in $`SO(3)`$ over the Wolf space $`\stackrel{~}{G}r_4(^6)`$, and this latter manifold is isometric to $`Gr_2(^4)`$. Thus again the geometry of 3-Sasakian and of Sasakian ($`\eta `$)-Einstein submanifolds $`\nu ^1(0)P^4`$ is obtained from the same starting point as Proposition 5.1. Besides this kind of geometry of submanifolds, the Sasakian-Einstein level sets $`\mu ^1(0)`$ share with some of the 3-Sasakian level sets $`\nu ^1(0)`$ a common expression of their Poincaré polynomials. We have in fact: ###### Proposition 5.2. The Poincaré polynomial of $`\mu ^1(0)P^n`$ is given by: $$Poin_{\mu ^1(0)}(t)=\underset{i=0}{\overset{[\frac{n1}{2}]}{}}(t^{4i}+t^{4n34i})$$ This is obtained by the Gysin sequence of the $`S^1`$-bundle $`\mu ^1(0)Gr_2(^{n+1})`$, where the connecting homomorphism $$H^p(Gr_2(^{n+1}))H^{p+2}(Gr_2(^{n+1}))$$ is given by the wedge product with the Kähler form of the Grassmannian. Since this wedge product is injective up to the middle real dimension $`p+2=2n2`$ (cf. , lemma 3.1, for the similar 3-Sasakian situation), the Gysin sequence reduces to a series of short exact sequences finishing with $$H^{2n4}(Gr_2(^{n+1}))H^{2n2}(Gr_2(^{n+1}))H^{2n2}(\mu ^1(0)).$$ This allows to compute the Betti numbers of $`(\mu ^1(0))`$ by differences of consecutive even Betti numbers in $`Gr_2(^{n+1})`$. The Poincaré polynomial of $`Gr_2(^{n+1})`$ is well known (see, for example, , p. 292) and by writing it as: (5.1) $$Poin_{Gr_2(^{n+1})}=(1+t^2+\mathrm{}+t^{2n2})(1+t^4+t^{4m4}),(n+1=2m)$$ (5.2) $$Poin_{Gr_2(^{n+1})}=(1+t^4+\mathrm{}+t^{4m4})(1+t^2+t^{2n}),(n+1=2m+1)$$ the conclusion is easily obtained. The following table gives the Poincaré polynomial of $`\mu ^1(0)P^n`$ for low values of $`n`$ The Poincaré polynomial of $`\mu ^1(0)`$ can be compared with that of $`\nu ^1(0)`$, computed as for the homogeneous 3-Sasakian manifold $`SO(n+1)/(SO(n3)\times Sp(1))`$. The latter has two different expressions, according to whether $`n+1`$ is even or odd (see or , p. 28). For odd values of $`n+1=2k+3`$ this expression is: $$Poin_{\nu ^1(0)}(t)=\underset{i=0}{\overset{k1}{}}(t^{4i}+t^{8k14i}),$$ and taking account of the dimensions, this is the same formula given in Proposition 5.2 for $`Poin_{\mu ^1(0)}(t)`$.
warning/0001/hep-ph0001239.html
ar5iv
text
# 1 Introduction ## 1 Introduction The understanding of hadron properties in the vacuum as well as in hot or dense matter is one of the central tasks of present-day nuclear physics. In principle, all properties of strongly interacting particles should be derived from QCD. However, at least in the low-energy regime, where perturbation theory is not applicable, this is presently limited to a rather small number of observables which can be studied on the lattice, while more complex processes have to be described within model calculations. So far the best descriptions of hadronic spectra, decays and scattering processes are obtained within phenomenological hadronic models. For instance the pion electromagnetic form factor in the time-like region can be reproduced rather well within a simple vector dominance model with a dressed $`\rho `$-meson which is constructed by coupling a bare $`\rho `$-meson to a two-pion intermediate state . Models of this type have been successfully extended to investigate medium modifications of vector mesons and to calculate dilepton production rates in hot and dense hadronic matter . In this situation one might ask how the phenomenologially successful hadronic models emerge from the underlying quark structure and the symmetry properties of QCD. Since this question cannot be answered at present from first principles it has to be addressed within quark models. For light hadrons chiral symmetry and its spontaneous breaking in the physical vacuum through instantons plays the decisive role in describing the two-point correlators with confinement being much less important. This feature is captured by the Nambu–Jona-Lasinio(NJL) model in which the four-fermion interactions can be viewed as being induced by instantons. The study of hadrons within the NJL model has of course a long history. In fact, mesons of various quantum numbers have already been discussed in the original papers by Nambu and Jona-Lasinio and by many authors thereafter (for reviews see ). Most of these works correspond to a leading-order approximation in $`1/N_c`$, the inverse number of colors. In this scheme quark masses are calculated in mean-field approximation and mesons are constructed as correlated quark-antiquark states. With the appropriate choice of parameters chiral symmetry, which is an (approximate) symmetry of the model Lagrangian, is spontaneously broken in the vacuum and pions emerge as (nearly) massless Goldstone bosons. While this is clearly one of the successes of the model, the description of other mesons is more problematic. One reason is the fact that the NJL model does not confine quarks. As a consequence a meson can decay into free constituent quarks if its mass is larger than twice the constituent quark mass $`m`$. Hence, for a typical value of $`m`$ 300 MeV, the $`\rho `$-meson with a mass of 770 MeV, for instance, would be unstable against decay into quarks. On the other hand the physical decay channel of the $`\rho `$-meson into two pions is not included in the standard approximation. Hence, even if a large constituent quark mass is chosen in order to suppress the unphysical decays into quarks, one obtains a poor description of the $`\rho `$-meson propagator and related observables, like the pion electromagnetic form factor. This and other reasons have motivated several authors to go beyond the standard approximation scheme and to include mesonic fluctuations. In Ref. a quark-antiquark $`\rho `$-meson is coupled via a quark triangle to a two-pion state. Also higher-order corrections to the quark self-energy and to the quark condensate have been investigated. However, as the most important feature of the NJL model is chiral symmetry, one should use an approximation scheme which conserves the symmetry properties, to ensure the existence of massless Goldstone bosons. Two slightly different symmetry conserving approximation schemes are discussed in Refs. and . The authors of Ref. use an expansion in $`1/N_c`$ in order to include mesonic fluctuations and calculate the changes of $`f_\pi `$ and of the quark condensate $`\overline{\psi }\psi `$ in a low-momentum expansion for the incoming state. In Ref. a correction term to the quark self-energy is included in the gap equation. In a perturbative scheme this term would be of order $`1/N_c`$ but, as the modified gap equation is solved self-consistently, arbitrary orders in $`1/N_c`$ are generated. The authors find a consistent scheme to describe mesons and show the validity of the Goldstone theorem and the Goldberger-Treiman relation in that scheme. Based on this model various authors have investigated the effect of meson-loop corrections on the pion electromagnetic form factor and on $`\pi `$-$`\pi `$ scattering in the vector and the scalar channel . However, since the numerical calculation of the multi-loop diagrams, which have to be evaluated, is rather involved, in these references the exact expressions are approximated by low-momentum expansions. In the present paper we calculate the $`\rho `$-meson self energy in a strict expansion up to next-to-leading order without any approximations. Within the same scheme we have recently studied the influence of mesonic fluctuations on the pion propagator . This was mainly motivated by recent works by Kleinert and Van den Bossche , who claim that chiral symmetry is not spontaneously broken in the NJL model as a result of strong mesonic fluctuations. In Ref. we argue that because of the non-renormalizability of the NJL model new divergences and hence new cutoff parameters emerge if one includes meson loops. Following Refs. and we regularize the meson loops by an independent cutoff parameter $`\mathrm{\Lambda }_M`$. The results are, of course, strongly dependent on this parameter. Whereas for moderate values of $`\mathrm{\Lambda }_M`$ the pion properties change only quantitatively strong instabilities are encountered for larger values of $`\mathrm{\Lambda }_M`$, which might be a hint for an instability of the spontaneously broken vacuum state. In Ref. we restricted ourselves to calculate the pion mass $`m_\pi `$, the pion decay constant $`f_\pi `$ and the quark condensate $`\overline{\psi }\psi `$. Since these quantities can already be fitted without meson-loop corrections (i.e. $`\mathrm{\Lambda }_M`$ = 0) one has to look at other observables in order to fix $`\mathrm{\Lambda }_M`$. Obviously the spectral function of the $`\rho `$-meson is particularly suited, as it cannot be described realistically without taking into account pion loops. In the present article all parameters of the NJL model will be fixed by fitting $`\rho `$-meson properties simultaneously with $`f_\pi `$ and $`\overline{\psi }\psi `$. The important result is that such a fit can indeed be achieved with a constituent quark mass which is large enough such that the unphysical $`q\overline{q}`$-threshold opens above the $`\rho `$-meson peak. Since the constituent quark mass is not an independent input parameter this was not clear a priori. The paper is organized as follows. In Sec. 2 we present the scheme for describing mesons in next-to-leading order in $`1/N_c`$. The consistency of this scheme with the Goldstone theorem and with the Gell-Mann Oakes Renner relation will be shown in Sec. 3. In Sec. 4 we perform a low-momentum expansion of the effective meson vertices and discuss the relation to hadronic models. The numerical results will be presented in Sec. 5. For several values of $`\mathrm{\Lambda }_M`$ we fix a subset of the model parameters by fitting quantities in the pion sector. Then we determine the remaining parameters, in particular $`\mathrm{\Lambda }_M`$, in the $`\rho `$-meson sector. Finally, conclusions are drawn in Sec. 6. ## 2 The NJL model in leading order and next-to-leading order in $`1/N_c`$ We consider the following NJL-type Lagrangian: $$=\overline{\psi }(i/m_0)\psi +g_s[(\overline{\psi }\psi )^2+(\overline{\psi }i\gamma _5\stackrel{}{\tau }\psi )^2]g_v[(\overline{\psi }\gamma ^\mu \stackrel{}{\tau }\psi )^2+(\overline{\psi }\gamma ^\mu \gamma _5\stackrel{}{\tau }\psi )^2].$$ (1) Here $`\psi `$ is a quark field with $`N_f`$ = 2 flavors and $`N_c`$ colors, while $`g_s`$ and $`g_v`$ are dimensionful coupling constants of the order $`1/N_c`$. In order to establish the counting scheme, the number of colors has been treated as variable, but all numerical calculations will be done with the physical value, $`N_c`$ = 3. In most publications the NJL model has been treated in leading order in $`1/N_c`$. In terms of many-body theory this corresponds to a (Bogoliubov) Hartree approximation for the quark propagator and to a random phase approximation (RPA) for describing mesonic excitations. Diagrammatically this is shown in Fig. 1. The selfconsistent solution of the Dyson equation shown in the upper part of Fig. 1 leads to a momentum independent quark self energy and therefore only gives a correction to the constituent quark mass: $$m=m_0+\mathrm{\hspace{0.33em}2}ig_s4N_cN_f\frac{d^4p}{(2\pi )^4}\frac{m}{p^2m^2+iϵ}.$$ (2) Usually one refers to this equation as the gap equation. For sufficiently large couplings $`g_s`$ it allows for a finite constituent quark mass $`m`$ even in the chiral limit, i.e. for $`m_0`$ = 0. In the mean-field approximation this solution minimizes the ground-state energy. Since $`g_s`$ is of order $`1/N_c`$ the constituent quark mass $`m`$, and hence the quark propagator $`S(p)=(p/m)^1`$ are of the order unity. Mesons are described via a Bethe-Salpeter equation, as shown in the lower part of Fig. 1. Since this is all standard we only list the results, which are needed later on. First we define the quark-antiquark polarization functions $$\mathrm{\Pi }_M(q)=i\frac{d^4p}{(2\pi )^4}\mathrm{Tr}[\mathrm{\Gamma }_MiS(p+\frac{q}{2})\mathrm{\Gamma }_MiS(p\frac{q}{2})],$$ (3) with $`M=\sigma ,\pi ,\rho ,a_1`$ and $`\mathrm{\Gamma }_\sigma =11`$, $`\mathrm{\Gamma }_\pi ^k=i\gamma _5\tau ^k`$, $`\mathrm{\Gamma }_\rho ^{\mu k}=\gamma ^\mu \tau ^k`$ and $`\mathrm{\Gamma }_{a_1}^{\mu k}=\gamma ^\mu \gamma _5\tau ^k`$. Here “$`\mathrm{Tr}`$” denotes a trace in color, flavor and Dirac space. $`\mathrm{\Pi }_M`$ is diagrammatically shown in Fig. 3. Iterating the scalar (pseudoscalar) part of the four-fermion interaction one obtains for the sigma meson (pion): $$D_\sigma (q)=\frac{2g_s}{12g_s\mathrm{\Pi }_\sigma (q)},D_\pi ^{ab}(q)D_\pi (q)\delta _{ab}=\frac{2g_s}{12g_s\mathrm{\Pi }_\pi (q)}\delta _{ab}.$$ (4) Here $`a`$ and $`b`$ are isospin indices and we have used the notation $`\mathrm{\Pi }_\pi ^{ab}(q)\mathrm{\Pi }_\pi (q)\delta _{ab}`$. In the vector channel this can be done in a similar way. Using the transverse structure of the polarization loop in the vector channel, $$\mathrm{\Pi }_\rho ^{\mu \nu ,ab}(q)=\mathrm{\Pi }_\rho (q)T^{\mu \nu }\delta _{ab};T^{\mu \nu }=(g^{\mu \nu }+\frac{q^\mu q^\nu }{q^2}),$$ (5) one obtains for the $`\rho `$-meson $$D_\rho ^{\mu \nu ,ab}(q)D_\rho (q)T^{\mu \nu }\delta _{ab}=\frac{2g_v}{12g_v\mathrm{\Pi }_\rho (q)}T^{\mu \nu }\delta _{ab}.$$ (6) Analogously, the $`a_1`$ can be constructed from the transverse part of the axial polarization function $`\mathrm{\Pi }_{a_1}`$. As discussed e.g. in Ref. $`\mathrm{\Pi }_{a_1}^{\mu \nu }`$ also contains a longitudinal part which contributes to the pion. Although there is no conceptional problem to include this mixing we will neglect it in the present paper in order to keep the structure of the model as simple as possible. It follows from Eqs. (3) - (6) that the functions $`D_M(q)`$ are of order $`1/N_c`$. Their explicit forms are given in App. B. For simplicity we will call them “propagators”, although strictly speaking, they should be interpreted as the product of a renormalized meson propagator with a squared quark-meson coupling constant. The latter is given by the inverse residue of the function $`D_M(q)`$, while the pole position determines the meson mass: $$D_M^1(q)|_{q^2=m_M^{2(0)}}=\mathrm{\hspace{0.33em}0},g_{Mqq}^{2(0)}=\frac{d\mathrm{\Pi }_M(q)}{dq^2}|_{q^2=m_M^{2(0)}}.$$ (7) We have used the superscript $`(0)`$ to indicate that $`m_M^{2(0)}`$ and $`g_{Mqq}^{2(0)}`$ are leading-order quantities in $`1/N_c`$. One easily verifies that they are of order unity and $`1/\sqrt{N_c}`$, respectively. With the help of the gap equation, Eq. (2), one can show that the pion is massless in the chiral limit, demonstrating the consistency of the scheme with chiral symmetry . We now turn to the corrections in next-to-leading order in $`1/N_c`$. The correction terms to the quark self-energy are shown in Fig. 2. In these diagrams the quark lines and the double lines correspond to quark propagators in the Hartree approximation (order unity) and to meson propagators in the RPA (order $`1/N_c`$), respectively. Both diagrams are therefore of order $`1/N_c`$. The $`1/N_c`$-corrected mesonic polarization diagrams read $$\stackrel{~}{\mathrm{\Pi }}_M(q)=\mathrm{\Pi }_M(q)+\underset{k=a,b,c,d}{}\delta \mathrm{\Pi }_M^{(k)}(q).$$ (8) The four correction terms $`\delta \mathrm{\Pi }_M^{(a)}`$ \- $`\delta \mathrm{\Pi }_M^{(d)}`$ together with the leading-order term $`\mathrm{\Pi }_M`$ are shown in Fig. 3. Again the lines in this figure correspond to Hartree quarks and RPA mesons. Besides the RPA meson propagators the main building blocks are quark triangle and box diagrams, which are shown in Fig. 4. The triangle diagrams entering into $`\delta \mathrm{\Pi }_M^{(a)}`$ and $`\delta \mathrm{\Pi }_M^{(d)}`$ can be interpreted as effective three-meson vertices. For external mesons $`M_1`$, $`M_2`$ and $`M_3`$ they are given by $`i\mathrm{\Gamma }_{M_1,M_2,M_3}(q,p)`$ $`=`$ $`{\displaystyle }{\displaystyle \frac{d^4k}{(2\pi )^4}}\{\mathrm{Tr}[\mathrm{\Gamma }_{M_1}iS(k)\mathrm{\Gamma }_{M_2}iS(kp)\mathrm{\Gamma }_{M_3}iS(k+q)]`$ (9) $`+\mathrm{Tr}[\mathrm{\Gamma }_{M_1}iS(kq)\mathrm{\Gamma }_{M_3}iS(k+p)\mathrm{\Gamma }_{M_2}iS(k)]\},`$ with the operators $`\mathrm{\Gamma }_M`$ as defined below Eq. (3). We have summed over both possible orientations of the quark loop. The quark box diagrams are effective four-meson vertices and are needed for the evaluation of $`\delta \mathrm{\Pi }_M^{(b)}`$ and $`\delta \mathrm{\Pi }_M^{(c)}`$. If one again sums over both orientations of the quark loop they are given by $`i\mathrm{\Gamma }_{M_1,M_2,M_3,M_4}(p_1,p_2,p_3)`$ (10) $`=`$ $`{\displaystyle }{\displaystyle \frac{d^4k}{(2\pi )^4}}(\mathrm{Tr}[\mathrm{\Gamma }_{M_1}iS(k)\mathrm{\Gamma }_{M_2}iS(kp_2)\mathrm{\Gamma }_{M_3}iS(kp_2p_3)\mathrm{\Gamma }_{M_4}iS(k+p_1)]`$ $`+\mathrm{Tr}[\mathrm{\Gamma }_{M_1}iS(kp_1)\mathrm{\Gamma }_{M_4}iS(k+p_2+p_3)\mathrm{\Gamma }_{M_3}iS(k+p_2)\mathrm{\Gamma }_{M_2}iS(k)]).`$ The polarization diagrams $`\delta \mathrm{\Pi }_M^{(a)}`$ \- $`\delta \mathrm{\Pi }_M^{(d)}`$ contain several loops. However, with the help of the definitions given above they can be written in a relatively compact form: $`\delta \mathrm{\Pi }_M^{(a)}(q)`$ $`=`$ $`{\displaystyle \frac{i}{2}}{\displaystyle \frac{d^4p}{(2\pi )^4}\underset{M_1M_2}{}\mathrm{\Gamma }_{M,M_1,M_2}(q,p)D_{M_1}(p)\mathrm{\Gamma }_{M,M_1,M_2}(q,p)D_{M_2}(pq)},`$ $`\delta \mathrm{\Pi }_M^{(b)}(q)`$ $`=`$ $`i{\displaystyle \frac{d^4p}{(2\pi )^4}\underset{M_1}{}\mathrm{\Gamma }_{M,M_1,M_1,M}(q,p,p)D_{M_1}(p)},`$ $`\delta \mathrm{\Pi }_M^{(c)}(q)`$ $`=`$ $`{\displaystyle \frac{i}{2}}{\displaystyle \frac{d^4p}{(2\pi )^4}\underset{M_1}{}\mathrm{\Gamma }_{M,M_1,M,M_1}(q,p,q)D_{M_1}(p)},`$ $`\delta \mathrm{\Pi }_M^{(d)}(q)`$ $`=`$ $`{\displaystyle \frac{i}{2}}\mathrm{\Gamma }_{M,M,\sigma }(q,q)D_\sigma (0){\displaystyle \frac{d^4p}{(2\pi )^4}\underset{M_1}{}\mathrm{\Gamma }_{M_1,M_1,\sigma }(p,p)D_{M_1}(p)}.`$ (11) Note that for $`\delta \mathrm{\Pi }_M^{(c)}`$ and $`\delta \mathrm{\Pi }_M^{(d)}`$ one has to include a symmetry factor of $`1/2`$, because otherwise the sum over the two orientations of the quark propagators, which is contained in the definitions of the quark triangle vertex (Eq. (9)) and of the quark box vertex (Eq. (10)) would lead to double counting. Similarly in $`\delta \mathrm{\Pi }_M^{(a)}`$ we had to correct for the fact that the exchange of $`M_1`$ and $`M_2`$ leads to identical diagrams. For the evaluation of Eq. (11) we have to proceed in two steps. In the first step we calculate the intermediate RPA meson-propagators. Simultaneously we can calculate the quark triangles and box diagrams. One is then left with a meson loop which has to be evaluated in a second step. The various sums are, in principle, over all quantum numbers of the intermediate mesons. Of course, many combinations vanish, e.g. because of isospin conservation. In fact, the expression for $`\delta \mathrm{\Pi }_M^{(d)}(q)`$ should contain a sum over the quantum numbers of both intermediate mesons. However, one can easily verify that the meson which connects the two quark triangles has to be a sigma meson in order to give a non-vanishing contribution. As we have mentioned earlier, the main focus of the present paper is on the $`\rho `$-meson. In this case the most important contribution comes from the two-pion intermediate state in diagram $`\delta \mathrm{\Pi }_M^{(a)}`$. Other contributions to this diagram, i.e $`\pi a_1`$, $`\rho \sigma `$, $`\rho \rho `$ and $`a_1a_1`$ intermediate states, are much less important since the corresponding decay channels open far above the $`\rho `$-meson mass and \- in the NJL model - also above the unphysical two-quark threshold. Hence, from a purely phenomenological point of view, it should be sufficient to restrict the sums in Eq. (11) to intermediate pions. However, in order to stay consistent with chiral symmetry, we have to include intermediate sigma mesons as well. On the other hand, vector- and axial-vector mesons can be neglected without violating chiral symmetry. Since this leads to an appreciable simplification of the numerics we have restricted the intermediate degrees of freedom to scalar and pseudoscalar mesons in the present paper. Of course, in order to describe a $`\rho `$-meson, we have to take vector couplings at the external vertices of the diagrams shown in Fig. 3. In analogy to Eqs. (4) and (6) the corrected meson propagators are given by $$\stackrel{~}{D}_M(q)=\frac{2g_M}{12g_M\stackrel{~}{\mathrm{\Pi }}_M(q)},$$ (12) with $`g_M=g_s`$ for $`\sigma `$ and $`\pi `$ and $`g_M=g_v`$ for $`\rho `$ and $`a_1`$. Since our scheme corresponds to a $`1/N_c`$-expansion of the polarization diagrams and hence of the inverse meson propagators, the $`\stackrel{~}{D}_M`$ contains terms of arbitrary orders in $`1/N_c`$. The corrected meson masses are again defined by the pole positions of the propagators: $$\stackrel{~}{D}_M^1(q)|_{q^2=m_M^2}=\mathrm{\hspace{0.33em}0}.$$ (13) As we will show in the next section our scheme is consistent with the Goldstone theorem, i.e. in the chiral limit it leads to massless pions. Note, however, that because of its implicit definition $`m_M`$ contains terms of arbitrary orders in $`1/N_c`$, although we start from a strict expansion of the inverse meson propagator up to next-to-leading order. This will be important in the context of the Gell-Mann Oakes Renner relation. Finally we should comment on Ref. , where the authors introduce $`1/N_c`$-corrections in a slightly different way. In that scheme a correction term to the quark self-energy is self-consistently included in the gap equation. This leads to a modified quark mass as compared to the ’Hartree mass’. Mesons are described by iterating the polarization diagrams $`\mathrm{\Pi }_M`$, $`\delta \mathrm{\Pi }_M^{(a)}`$, $`\delta \mathrm{\Pi }_M^{(b)}`$ and $`\delta \mathrm{\Pi }_M^{(c)}`$, but using the modified quarks for constructing RPA mesons as well as quark triangle- and box diagrams. Since the $`1/N_c`$-correction terms are iterated in the gap equation, all diagrams contain terms of arbitrary orders in $`1/N_c`$. However, if one performs a strict $`1/N_c`$-expansion of the mesonic polarization diagrams one exactly recovers our scheme of describing mesons . Both schemes, the one introduced in Ref. and the strict $`1/N_c`$-expansion, are consistent with chiral symmetry. In particular, they lead to massless pions in the chiral limit. However, because of the modified gap equation, in the scheme of Ref. the RPA pions are not massless in the chiral limit, but tachyonic. Therefore this model is not very well suited for calculating the $`\rho `$-meson self-energy, which is mainly determined by intermediate RPA pions. For this reason we prefer the strict $`1/N_c`$-expansion scheme, in which both RPA- and $`1/N_c`$-corrected pions are massless in the chiral limit. Still there remains the problem that $`m_\pi m_\pi ^{(0)}`$, if we go away from the chiral limit, but at least for our final parameter set the difference is only about 10%. This will be discussed in more detail in Sec. 5. ## 3 Consistency with chiral symmetry Before discussing the numerical details we wish to show that the scheme introduced in the previous section is consistent with chiral symmetry. We begin with the Goldstone theorem and then show the consistency of the scheme with the Gell-Mann Oakes Renner relation. This is not an entirely academic exercise. Since most of the integrals which have to be evaluated are divergent and hence must be regularized one has to ensure that the various symmetry relations are not destroyed by the regularization. To this end, it is important to know how these relations formally emerge. The validity of the Goldstone theorem in our scheme has already been proven by Dmitrašinović et al. and we only summarize the main steps which are relevant for our later discussion. One has to show that, in the chiral limit, the inverse pion propagator vanishes at zero momentum, $$2g_s\stackrel{~}{\mathrm{\Pi }}_\pi (0)=\mathrm{\hspace{0.33em}1}\mathrm{for}m_0=\mathrm{\hspace{0.33em}0}.$$ (14) As before we use the notation $`\stackrel{~}{\mathrm{\Pi }}_\pi ^{ab}=\delta _{ab}\stackrel{~}{\mathrm{\Pi }}_\pi `$. Restricting the calculation to the chiral limit and to zero momentum simplifies the expressions considerably and Eq. (14) can be proven analytically. Since the Goldstone theorem is fulfilled in leading order, i.e. $`2g_s\mathrm{\Pi }_\pi (0)`$ = 1 for $`m_0`$ = 0, we only need to show that the contributions of the correction terms add up to zero: $$\underset{k=a,b,c,d}{}\delta \mathrm{\Pi }_\pi ^{(k)}(0)=\mathrm{\hspace{0.33em}0}\mathrm{for}m_0=\mathrm{\hspace{0.33em}0}.$$ (15) Let us begin with diagram $`\delta \mathrm{\Pi }_\pi ^{(a)}`$. As mentioned above, we neglect the $`\rho `$ and $`a_1`$ subspace for intermediate mesons. Then one can easily see that the external pion can only couple to a $`\pi \sigma `$ intermediate state. Evaluating the trace in Eq. (9) for zero external momentum one gets for the corresponding triangle diagram: $$\mathrm{\Gamma }_{\pi ,\pi ,\sigma }^{ab}(0,p)=\delta _{ab}4N_cN_f2mI(p),$$ (16) with $`a`$ and $`b`$ being isospin indices and the elementary integral $$I(p)=\frac{d^4k}{(2\pi )^4}\frac{1}{(k^2m^2+i\epsilon )((k+p)^2m^2+i\epsilon )}.$$ (17) Inserting this into Eq. (11) we find $$\delta \mathrm{\Pi }_\pi ^{(a)ab}(0)=i\delta _{ab}\frac{d^4p}{(2\pi )^4}(4N_cN_fI(p))^24m^2D_\sigma (p)D_\pi (p).$$ (18) Now the essential step is to realize that the product of the RPA sigma- and pion propagators can be converted into a difference , $$D_\sigma (p)D_\pi (p)=i\frac{D_\sigma (p)D_\pi (p)}{4N_cN_f2m^2I(p)},$$ (19) to finally obtain $`\delta \mathrm{\Pi }_\pi ^{(a)ab}(0)=\delta _{ab}\mathrm{\hspace{0.33em}4}N_cN_f{\displaystyle \frac{d^4p}{(2\pi )^4}2I(p)\left\{D_\sigma (p)D_\pi (p)\right\}}.`$ (20) The next two diagrams can be evaluated straightforwardly. One finds: $`\delta \mathrm{\Pi }_\pi ^{(b)ab}(0)`$ $`=`$ $`\delta _{ab}\mathrm{\hspace{0.33em}4}N_cN_f{\displaystyle }{\displaystyle \frac{d^4p}{(2\pi )^4}}\{D_\sigma (p)(I(p)+I(0)(p^24m^2)K(p))`$ $`+D_\pi (p)(3I(p)+\mathrm{\hspace{0.33em}3}I(0)\mathrm{\hspace{0.33em}\hspace{0.33em}3}p^2K(p))\},`$ $`\delta \mathrm{\Pi }_\pi ^{(c)ab}(0)`$ $`=`$ $`\delta _{ab}\mathrm{\hspace{0.33em}4}N_cN_f{\displaystyle \frac{d^4p}{(2\pi )^4}I(p)\left\{D_\sigma (p)D_\pi (p)\right\}}.`$ (21) The elementary integral $`K(p)`$ is of the same type as the integral $`I(p)`$ and is defined in App. A. Finally we have to calculate $`\delta \mathrm{\Pi }_\pi ^{(d)}(0)`$. According to Eq. (11), it can be written in the form $$\delta \mathrm{\Pi }_\pi ^{(d)ab}(0)=i\mathrm{\Gamma }_{\pi ,\pi ,\sigma }^{ab}(0,0)D_\sigma (0)\mathrm{\Delta },$$ (22) with $`\mathrm{\Delta }`$ $`=`$ $`{\displaystyle \frac{1}{2}}{\displaystyle \frac{d^4p}{(2\pi )^4}\underset{M}{}(iD_M(p))(i\mathrm{\Gamma }_{M,M,\sigma }(p,p))}`$ (23) $`=`$ $`4N_cN_fm{\displaystyle }{\displaystyle \frac{d^4p}{(2\pi )^4}}\{D_\sigma (p)(2I(p)+I(0)(p^24m^2)K(p))`$ $`+D_\pi (p)(\mathrm{\hspace{0.33em}3}I(0)\mathrm{\hspace{0.33em}3}p^2K(p))\}.`$ Evaluating $`D_\sigma (0)`$ in the chiral limit and comparing the result with Eq. (16) one finds that the product of the first two factors in Eq. (22) is simply $`\delta _{ab}/m`$, i.e. one gets $$\delta \mathrm{\Pi }_\pi ^{(d)ab}(0)=\delta _{ab}\frac{\mathrm{\Delta }}{m}.$$ (24) With these results one can easily check that Eq. (15) indeed holds in our scheme. Of course, most of the integrals we have to deal with are divergent and have to be regularized. Therefore one has to make sure that all steps which lead to Eq. (15) remain valid in the regularized model. One important observation is that the cancellations occur already on the level of the $`p`$-integrand, i.e. before performing the meson-loop integral. This means that there is no restriction on the regularization of this loop. We also do not need to perform the various quark loop integrals explicitly but can make use of several relations between them. For instance, in order to arrive at Eq. (24) we need the similar structure of quark triangle $`\mathrm{\Gamma }_{\pi ,\pi ,\sigma }(0,0)`$ and the inverse RPA propagator $`D_\sigma (0)^1`$. Therefore all quark loops, i.e. RPA polarizations, triangles and box diagrams should be consistently regularized within the same scheme, whereas the meson loops can be regularized independently. Going away from the chiral limit the pion gets a finite mass. To lowest order in the current quark mass it is given by the Gell-Mann Oakes Renner (GOR) relation, $$m_\pi ^2f_\pi ^2=m_0\overline{\psi }\psi .$$ (25) Our scheme is consistent with this relation up to next-to-leading order in $`1/N_c`$. Expanding both sides of Eq. (25) we get $$m_\pi ^{2(0)}f_\pi ^{2(0)}+m_\pi ^{2(0)}\delta f_\pi ^2+\delta m_\pi ^2f_\pi ^{2(0)}=m_0\left(\overline{\psi }\psi ^{(0)}+\delta \overline{\psi }\psi \right).$$ (26) Here $`m_\pi ^{2(0)}`$ denotes the leading-order contribution to the squared pion mass, while $`\delta m_\pi ^2`$ is the next-to-leading order contribution. For $`f_\pi ^2`$ and $`\overline{\psi }\psi `$ we introduce similar notations. Since the GOR relation holds only in lowest order in $`m_0`$, Eq. (26) corresponds to a double expansion: $`m_\pi ^2`$ has to be calculated in linear order in $`m_0`$, $`f_\pi ^2`$ and $`\overline{\psi }\psi `$ in the chiral limit. The quark condensate is given by $$\overline{\psi }\psi =i\frac{d^4p}{(2\pi )^4}Tr\stackrel{~}{S}(p),$$ (27) with the $`1/N_c`$-corrected quark propagator $$\stackrel{~}{S}(p)=S(p)+S(p)(\mathrm{\Sigma }^{(a)}+\mathrm{\Sigma }^{(b)}(p))S(p).$$ (28) Here $`S(p)`$ is the Hartree propagator and $`\mathrm{\Sigma }^{(a)}`$ and $`\mathrm{\Sigma }^{(b)}`$ are the self energy corrections shown in Fig. 2. Since we are interested in a strict $`1/N_c`$ expansion, it should not be iterated. The r.h.s. of Eq. (27) can be easily evaluated and one obtains $$\overline{\psi }\psi ^{(0)}+\delta \overline{\psi }\psi =\frac{mm_0}{2g_s}\frac{D_\sigma (0)\mathrm{\Delta }}{2g_s},$$ (29) with $`\mathrm{\Delta }`$ as defined in Eq. (23). The pion decay constant $`f_\pi `$ is calculated from the one-pion to vacuum axial vector matrix element. Basically this corresponds to evaluating the mesonic polarization diagrams, Fig. 3, coupled to an external axial current and to a pion. This leads to expressions similar to Eqs. (3) and (11), but with one external vertex equal to $`\gamma ^\mu \gamma _5\frac{\tau ^a}{2}`$, corresponding to the axial current, and the second external vertex equal to $`g_{\pi qq}i\gamma _5\tau ^b`$, corresponding to the pion. Here the $`1/N_c`$-corrected pion-quark coupling constant is defined as $$g_{\pi qq}^2=g_{\pi qq}^{2(0)}+\delta g_{\pi qq}^2=\frac{d\stackrel{~}{\mathrm{\Pi }}_\pi (q)}{dq^2}|_{q^2=m_\pi ^2},$$ (30) analogously to Eq. (7). Now we take the divergence of the axial current and then use the axial Ward identity $$\gamma _5p/=\mathrm{\hspace{0.33em}2}m\gamma _5+\gamma _5S^1(k+p)+S^1(k)\gamma _5$$ (31) to simplify the expressions . One finds: $$f_\pi =g_{\pi qq}\left(\frac{\stackrel{~}{\mathrm{\Pi }}_\pi (q)\stackrel{~}{\mathrm{\Pi }}_\pi (0)}{q^2}m+\frac{\mathrm{\Pi }_\pi (q)\mathrm{\Pi }_\pi (0)}{q^2}D_\sigma (0)\mathrm{\Delta }\right)|_{q^2=m_\pi ^2}.$$ (32) In the chiral limit, $`q^2=m_\pi ^20`$, Eqs. (7) and (30) can be employed to replace the difference ratios on the r.h.s. by pion-quark coupling constants. When we square this result and only keep the leading order and the next-to-leading order in $`1/N_c`$ to finally obtain: $$f_\pi ^{2(0)}+\delta f_\pi ^2=m^2g_{\pi qq}^{2(0)}+\left(m^2\delta g_{\pi qq}^2+\mathrm{\hspace{0.33em}2}mD_\sigma (0)\mathrm{\Delta }g_{\pi qq}^{2(0)}\right).$$ (33) For the pion mass we start from Eqs. (12) and (13) and expand the inverse pion propagator around $`q^2=0`$: $$1\mathrm{\hspace{0.33em}2}g_s\stackrel{~}{\mathrm{\Pi }}_\pi (0)\mathrm{\hspace{0.33em}2}g_s\left(\frac{d}{dq^2}\stackrel{~}{\mathrm{\Pi }}_\pi (q)\right)|_{q^2=0}m_\pi ^2+𝒪(m_\pi ^4)=\mathrm{\hspace{0.33em}0}.$$ (34) To find $`m_\pi ^2`$ in lowest non-vanishing order in $`m_0`$ we have to expand $`12g_s\stackrel{~}{\mathrm{\Pi }}_\pi (0)`$ up to linear order in $`m_0`$, while the derivative has to be calculated in the chiral limit, where it can be identified with the inverse squared pion-quark coupling constant, Eq. (30). The result can be written in the form $$m_\pi ^2=\frac{m_0}{m}\frac{g_{\pi qq}^2}{2g_s}\left(\mathrm{\hspace{0.17em}1}\frac{D_\sigma (0)\mathrm{\Delta }}{m}\right)+𝒪(m_0^2).$$ (35) Finally one has to expand this equation in powers of $`1/N_c`$. This amounts to expanding $`g_{\pi qq}^2`$, which is the only term in Eq. (35) which is not of a definite order in $`1/N_c`$. One gets: $$m_\pi ^{2(0)}+\delta m_\pi ^2=m_0\frac{m}{2g_s}\frac{g_{\pi qq}^{2(0)}}{m^2}m_0\frac{m}{2g_s}\frac{g_{\pi qq}^{2(0)}}{m^2}\left(g_{\pi qq}^{2(0)}\delta g_{\pi qq}^2+\frac{D_\sigma (0)\mathrm{\Delta }}{m}\right).$$ (36) It can be seen immediately that the leading-order term is exactly equal to $`m_0\overline{\psi }\psi ^{(0)}/f_\pi ^{2(0)}`$, as required by the GOR relation. Moreover, combining Eqs. (29) for $`m_0`$ = 0, (33) and (36) one finds that the GOR relation in next-to-leading order, Eq. (26), holds in our scheme. However, one should emphasize that this result is obtained by a strict $`1/N_c`$-expansion of the various properties which enter into the GOR relation and of the GOR relation itself. If one takes $`f_\pi `$ and $`m_\pi `$ as they result from Eqs. (32) and (35) and inserts them into the l.h.s. of Eq. (25) one will in general find deviations from the r.h.s. which are due to higher-order terms in $`1/N_c`$. In this sense one can take the violation of the GOR relation as a measure for the importance of these higher-order terms . ## 4 Relation to hadronic models This section discusses an approximation to our scheme which points out the relation to hadronic models. In order to suppress the quark effects in the present model, it is suggestive to assume that the constituent quark mass is very large as compared to the relevant meson momenta. This assumption leads to an effective zero momentum approximation, i.e. the quark vertices are taken at zero incoming momentum (“static limit”). In order to preserve chiral symmetry we then have to approximate the RPA-meson propagators consistently. This amounts to replacing the function $`I(p)`$ in the RPA polarization functions by $`I(0)`$. The latter can be related to the leading-order pion decay constant in this approximation. The RPA-meson propagators then take the form of free boson propagators: $$D_M^{(0)}(q)=\frac{g_{Mqq}^{2(0)}}{q^2m_M^{(0)2}},$$ (37) with $$m_\pi ^{2(0)}=\frac{m_0m}{2g_sf_\pi ^{2(0)}},m_\rho ^{2(0)}=\frac{3m^2}{4g_vf_\pi ^{2(0)}},m_\sigma ^{2(0)}=m_\pi ^{2(0)}+4m^2,m_{a_1}^{2(0)}=m_\rho ^{2(0)}+6m^2,$$ (38) and $$g_{\pi qq}^{2(0)}=g_{\sigma qq}^{2(0)}=\frac{2}{3}g_{\rho qq}^{2(0)}=\frac{2}{3}g_{a_1qq}^{2(0)}=\frac{m^2}{f_\pi ^{2(0)}}.$$ (39) Next we expand the quark triangles and box diagrams to first order in the external momenta. In this way one obtains a hadronic model with effective meson-meson coupling constants. Let us, for example, look at the $`\rho `$-meson self-energy which is generated as an approximation to the $`1/N_c`$-corrected NJL model. As the $`\rho \rho \sigma `$\- and the $`\rho \rho \sigma \sigma `$-vertex vanish in this approximation, we are left with a pion loop diagram, which is generated from the diagram shown in Fig. 2(a), and a pion tadpole diagram, which arises from the sum of the diagrams in Fig. 2(b) and (c). These are exactly the diagrams which are calculated in standard hadronic descriptions for the $`\rho `$-meson in vacuum, e.g. . In hadronic models one is also interested in the coupling of a bare $`\rho `$-meson to a photon, which is e.g. needed to calculate the pion electromagnetic form factor via vector-meson dominance. In the NJL model the bare $`\rho `$-meson corresponds to the RPA meson and its coupling to a photon is basically given by the RPA polarization loop in the vector channel. If we now perform the same low-momentum approximations as for the RPA-meson propagators we find that the vertex is given by $`i(e/g_{\rho \pi \pi })q^2g_{\mu \nu }`$ which exactly corresponds to the $`\gamma \rho `$ vertex in the vector dominance model of Kroll, Lee and Zumino . For the effective $`\rho \pi \pi `$ coupling constant we obtain $$g_{\rho \pi \pi }^2=6g_{\pi qq}^{2(0)}=6\frac{m^2}{f_\pi ^{(0)2}}.$$ (40) This result can be used to estimate the importance of quark effects. If we take the commonly used value of about 6 for $`g_{\rho \pi \pi }`$ in hadronic models we obtain for the constituent quark mass $$m\sqrt{6}f_\pi ^{(0)}.$$ (41) As will be seen in the next section the leading-order pion decay constant is larger but not much larger than the $`1/N_c`$-corrected quantity. If we take $`f_\pi ^{(0)}`$ 100-150 MeV, we find that the constituent quark mass should be about 250-350 MeV, which is obviously in contradiction to our original assumption of very heavy quarks . Thus we would expect that this approximation does not describe the full model very well. This point will be discussed in more detail in the next section. Similar to the $`\rho `$-meson one can perform approximations to the $`1/N_c`$-corrected self-energies of the other mesons. The authors of Ref. have used effective meson-meson coupling constants generated from this approximation to the NJL quark loops in a linear sigma model calculation for in-medium pion properties. It is nice to see that a consistent approximation to the $`1/N_c`$-corrected NJL model naturally generates an effective one loop approximation to the linear sigma model in the $`\pi \sigma `$ sector. A slightly different approximation to the NJL model has been performed in ref. , where instead of a low-momentum expansion, the vertices are evaluated for on-shell intermediate mesons. However, at least for processes dominated by intermediate pions this gives very similar results to those obtained with the low-momentum expansion. ## 5 Numerical results In this section we present numerical results for the self-energy of the $`\rho `$-meson and related quantities, such as the electromagnetic form factor of the pion and $`\pi `$-$`\pi `$ phase shifts in the vector channel. Before we begin with the explicit calculation we have to come back to the regularization. As discussed in Sec. 3, all quark loops, i.e. the RPA polarization diagrams, the quark triangles and the quark box diagrams should be regularized in the same way in order to preserve chiral symmetry. We use a Pauli-Villars-regularization with two regulator masses $`\sqrt{m^2+\mathrm{\Lambda }_q^2}`$ and $`\sqrt{m^2+2\mathrm{\Lambda }_q^2}`$. As before, $`m`$ denotes the constituent quark mass and $`\mathrm{\Lambda }_q`$ is the cutoff parameter. The regularization of the meson loop (integration over $`d^4p`$ in eq. (11)) is not constrained by chiral symmetry and independent from the quark loop regularization. For practical reasons we choose a three-dimensional cutoff $`\mathrm{\Lambda }_M`$ in momentum space. In order to get a well-defined result we work in the rest frame of the $`1/N_c`$-improved meson. The same regularization scheme was already used in Ref. . With the additional meson cutoff $`\mathrm{\Lambda }_M`$ there are five parameters: the current quark mass $`m_0`$, the two coupling constants $`g_s`$ and $`g_v`$, the quark-loop cutoff $`\mathrm{\Lambda }_q`$ and the meson-loop cutoff $`\mathrm{\Lambda }_M`$. For a given value of $`\mathrm{\Lambda }_M`$, the current quark mass $`m_0`$, the scalar coupling constant $`g_s`$ and the cutoff $`\mathrm{\Lambda }_q`$ can be fixed by fitting the pion mass $`m_\pi `$, the pion decay constant $`f_\pi `$ and the quark condensate $`\overline{\psi }\psi `$ to their empirical values. Then we can try to determine the two remaining parameters, i.e. the vector coupling constant $`g_v`$ and the meson cutoff $`\mathrm{\Lambda }_M`$, by fitting the pion electromagnetic form factor in the time-like region, which is related, via vector meson dominance, to the $`\rho `$-meson propagator (see below). Roughly speaking this amounts to fitting the $`\rho `$-meson mass and its width. The $`\rho `$-meson mass is most sensitive to the coupling constant $`g_v`$. Since in the present article we neglect possible $`\rho `$ and $`a_1`$ meson intermediate states, the value of $`g_v`$ does not influence the observables in the pion sector, $`m_\pi ,f\pi `$ and $`\overline{\psi }\psi `$. Hence for given $`\mathrm{\Lambda }_M`$ we can try to fit these observables together with the $`\rho `$-meson mass. Finally, we determine $`\mathrm{\Lambda }_M`$ by comparing the resulting pion electromagnetic form factor with the experimental data. However, this procedure contains a problem: For $`m_00`$ the corrected pion mass $`m_\pi `$ is always larger than the RPA pion mass $`m_\pi ^{(0)}`$. Hence, if we perform a fit for the $`1/N_c`$-improved pion mass, the masses of the RPA pions, which enter into the intermediate states, e.g. in diagram Fig. 3(a), are too small, i.e. the value of the threshold energy for the decay of the $`\rho `$-meson into two pions is too small. Therefore, since we are mainly interested in the $`\rho `$-meson in this article, we adjust $`m_\pi ^{(0)}`$, rather than $`m_\pi `$ to the physical pion mass. Here we make use of the fact that the pion mass is very sensitive to the current quark mass $`m_0`$ whereas the two other observables we want to fit, $`f_\pi `$ and $`\overline{\psi }\psi `$, depend only weakly on $`m_0`$. Five parameter sets (corresponding to five different meson cutoff values $`\mathrm{\Lambda }_M`$) are listed in Table 1, together with the constituent quark mass $`m`$, the values of $`m_\pi `$, $`f_\pi `$ and $`\overline{\psi }\psi `$ and the corresponding leading-order quantities. As outlined above, $`\mathrm{\Lambda }_q`$, $`m_0`$ and $`g_s`$ have been obtained by fitting $`m_\pi ^{(0)}`$, $`f_\pi `$ and $`\overline{\psi }\psi `$ to the empirical values. For the quark condensate this is not known very precisely, but the absolute value is probably below 2(260 MeV)<sup>3</sup>. (This corresponds roughly to the upper limit extracted in ref. from sum rules at a renormalization scale of 1 GeV. Recent lattice results are $`\overline{\psi }\psi `$ = -2($`(231\pm 4\pm 8\pm 6)`$ MeV)<sup>3</sup> .) It turns out that we can only stay below this limit, if the meson cutoff is not too large ($`\mathrm{\Lambda }_M`$ 700 MeV). This is related to the fact that the constituent quark mass and hence the absolute value of the leading-order condensate $`\overline{\psi }\psi ^{(0)}`$ increases with $`\mathrm{\Lambda }_M`$ if we keep $`f_\pi `$ constant. Table 1 also displays the ratio $`m_0\overline{\psi }\psi /m_\pi ^2f_\pi ^2`$, which would be equal to 1 if the GOR relation was exactly fulfilled. Note that for all parameter sets given in the table the deviations are less than 10% (for $`\mathrm{\Lambda }_M`$ 600 MeV even less than 3%), indicating that higher-order corrections in $`1/N_c`$ are small. We now turn to the $`\rho `$-meson channel. According to Eq. (8), the polarization function of the $`\rho `$-meson is a sum of the RPA polarization loop and the four $`1/N_c`$-correction terms, as shown in Fig. 3: $$\stackrel{~}{\mathrm{\Pi }}_\rho ^{\mu \nu ,ab}(q)=\mathrm{\Pi }_\rho ^{\mu \nu ,ab}(q)+\underset{k=a,b,c,d}{}\delta \mathrm{\Pi }_\rho ^{(k)\mu \nu ,ab}(q).$$ (42) Because of vector current conservation, the polarization function has to be transverse, i.e. $$q_\mu \stackrel{~}{\mathrm{\Pi }}_\rho ^{\mu \nu ,ab}(q)=q_\nu \stackrel{~}{\mathrm{\Pi }}_\rho ^{\mu \nu ,ab}(q)=\mathrm{\hspace{0.33em}0}.$$ (43) For our scheme it can be shown with the help of Ward identities that these relations hold, if we assume that the regularization preserves this property. This is the case for the Pauli-Villars regularization scheme, which was employed to regularize the RPA part $`\mathrm{\Pi }_\rho `$. Together with Lorentz covariance this leads to Eq. (5) for the tensor structure of $`\mathrm{\Pi }_\rho `$. On the other hand, since we use a three-dimensional sharp cutoff for the regularization of the meson loops, the $`1/N_c`$-correction terms are in general not transverse. However, as mentioned above, we work in the rest frame of the $`\rho `$-meson, i.e. $`\stackrel{}{q}`$ = 0. In this particular case Eq. (43) is not affected by the cutoff and the entire function $`\stackrel{~}{\mathrm{\Pi }}_\rho `$ can be written in the form of Eq. (5): $$\stackrel{~}{\mathrm{\Pi }}_\rho ^{\mu \nu ,ab}(q)=\stackrel{~}{\mathrm{\Pi }}_\rho (q)T^{\mu \nu }\delta _{ab}=\left(\mathrm{\Pi }_\rho (q)+\underset{k=a,b,c,d}{}\delta \mathrm{\Pi }_\rho ^{(k)}(q)\right)T^{\mu \nu }\delta _{ab},$$ (44) i.e. instead of evaluating all tensor components separately we only need to calculate the scalar functions $`\mathrm{\Pi }_\rho =1/3g_{\mu \nu }\mathrm{\Pi }_\rho ^{\mu \nu }`$ and $`\delta \mathrm{\Pi }_\rho ^{(k)}=1/3g_{\mu \nu }\delta \mathrm{\Pi }_\rho ^{(k)\mu \nu }`$. A second consequence of vector current conservation is, that the polarization function should vanish for $`q^2=0`$. For the $`1/N_c`$-correction terms this is violated by the sharp cutoff. We cure this problem by performing a subtraction: $$\underset{k=a,b,c,d}{}\delta \mathrm{\Pi }_\rho ^{(k)}(q)\underset{k=a,b,c,d}{}\left(\delta \mathrm{\Pi }_\rho ^{(k)}(q)\delta \mathrm{\Pi }_\rho ^{(k)}(0)\right).$$ (45) Note, however, that already at the RPA level a subtraction is required, although the RPA part is regularized by Pauli-Villars. This is due to a rather general problem which is discussed in detail in App. B. As mentioned above, we want to fix the remaining two model parameters, i.e. $`g_v`$ and $`\mathrm{\Lambda }_M`$, by fitting the pion electromagnetic form factor, $`F_\pi (q)`$, in the time-like region, which is dominated by the $`\rho `$-meson. The diagrams we include in our calculations are shown in Fig. 5. The two diagrams in the upper part correspond to the standard NJL description of the form factor if the full $`\rho `$-meson propagator (curly line) is replaced by the RPA one. Hence, the first improvement is the use of the $`1/N_c`$-corrected $`\rho `$-meson propagator in our model. Since, in the standard scheme, the photon couples to the $`\rho `$-meson via a quark-antiquark polarization loop, in our scheme we should also take into account the $`1/N_c`$-corrections to the polarization diagram for consistency. This leads to the diagrams in the lower part of Fig. 5. On the other hand the external pions are taken to be RPA pions (i.e. mass $`m_\pi ^{(0)}`$ and pion-quark-quark coupling constant $`g_{\pi qq}^{(0)}`$). This is more consistent with the fact that the $`\rho `$-meson is also dressed by RPA pions and, as discussed above, we have fitted $`m_\pi ^{(0)}`$ to the experimental value. Evaluating the diagrams of Fig. 5 we obtain for the form factor $$|F_\pi (q)|^2=\frac{1}{2}\left|g_{\pi qq}^{(0)2}f(p,p^{})\left(1\stackrel{~}{\mathrm{\Pi }}_\rho (q)\stackrel{~}{D}_\rho (q)\right)\right|^2,$$ (46) where $`p`$ and $`p^{}`$ are the four-momenta of the external pions and $`f(p,p^{})`$ is a scalar function appearing in the $`\pi \pi \rho `$-vertex function (see App. C). For the form factor it has to be evaluated for on-shell pions, i.e. $`p^2=p^2=m_\pi ^2`$ and $`pp^{}=q^2/2m_\pi ^2`$. From its explicit form, which is given in App. C, one can show that $`g_{\pi qq}^{(0)2}f(p,p^{})=\sqrt{2}`$ for $`q^2=0`$. Since the $`\rho `$-meson self energy $`\stackrel{~}{\mathrm{\Pi }}_\rho `$ vanishes at this point, we find $`F_\pi (0)=1`$, as it should be. The numerical results for $`|F_\pi |^2`$ as a function of the center-of-mass energy squared are displayed in the left panel of Fig. 6, together with the experimental data . The various curves correspond to different values of the meson cutoff $`\mathrm{\Lambda }_M`$. For the other parameters the values listed in Table 1 are taken. The vector coupling constant $`g_v`$ is chosen such that the maximum of the form factor is at the correct position. Roughly speaking this corresponds to fitting the mass of the (dressed) $`\rho `$-meson. We only show the results for $`\mathrm{\Lambda }_M`$ 500 MeV. For lower meson cutoffs the unphysical quark-antiquark decay threshold of the $`\rho `$-meson is below the maximum of the form factor and makes a comparison not very meaningful. For $`\mathrm{\Lambda }_M`$ = 500 MeV (dashed line) the quark-antiquark threshold is at $`s`$ = 0.63 GeV<sup>2</sup>, causing a cusp in the form factor slightly above the maximum. Because of the sub-threshold attraction in the $`\rho `$-meson channel the form factor drops very steeply below the cusp, leading to a poor description of the data above the maximum. In addition, also at the rising edge of the peak, where threshold effects are less important, we see that the width of the form factor is underestimated by the calculation, i.e. $`\mathrm{\Lambda }_M`$ is too small. On the other hand, a meson cutoff of 700 MeV (dotted line) is already too large. In particular, the height of the maximum is underestimated. With $`\mathrm{\Lambda }_M`$ = 600 MeV (solid line) we obtain the best description of the data. Since we assumed exact isospin symmetry in our model we can, of course, not reproduce the detailed structure of the form factor around 0.61 GeV<sup>2</sup>, which is due to $`\rho `$-$`\omega `$-mixing. The high-energy part above the peak is somewhat underestimated, mainly due to the $`q\overline{q}`$-threshold at $`s`$ = 0.80 GeV<sup>2</sup>. Probably the fit can be somewhat improved if we take a slightly larger meson cutoff, but we are not interested in fine-tuning here. In addition, it is to be expected that the inclusion of $`\rho `$\- and $`a_1`$ intermediate states will improve the high-energy behavior. A closely related quantity is the charge radius of the pion, which is defined as $$r_\pi ^2=\mathrm{\hspace{0.33em}6}\frac{dF_\pi }{dq^2}|_{q^2=0}.$$ (47) Results for $`\mathrm{\Lambda }_M`$ = 500, 600 and 700 MeV are listed in Table 1. All of them are close to the experimental value, $`r_\pi ^2^{1/2}`$ = (0.663 $`\pm `$ 0.006) fm . For $`\mathrm{\Lambda }_M`$ = 700 MeV we find perfect agreement. On the other hand, a simple pole ansatz for the form factor leads to $`r_\pi ^2^{1/2}=\sqrt{6}/m_\rho `$ = 0.63 fm , which is of the same quality as our results. Obviously, at $`q^2`$ = 0 we are not very sensitive to the details of the $`\rho `$-meson peak, while other effects which are not included in our model might start to play a role. Therefore the pion charge radius is certainly not very well suited for fixing our model parameters. One can also look at the $`\pi \pi `$-phase shifts in the vector-isovector channel. We include the diagrams shown in Fig. 7, i.e. the s-channel $`\rho `$-meson exchange and the direct $`\pi \pi `$-scattering via a quark box diagram. The latter has to be projected onto spin and isospin 1, which is a standard procedure. (For example, the analogous projection onto spin and isospin 0 can be found in Refs. .) Our results, together with the empirical data , are displayed in the right panel of Fig. 6. Since the main contribution comes from the s-channel $`\rho `$-meson exchange, they more or less confirm our findings for the form factor: below the $`\rho `$-meson peak the best fit of the data is obtained with $`\mathrm{\Lambda }_M`$ = 600 MeV while, at higher energies, where $`q\overline{q}`$-threshold effects start to play a role, the data are better described with $`\mathrm{\Lambda }_M`$ = 700 MeV. Finally, we wish to compare the results of the full $`1/N_c`$-corrected NJL model with the static limit, i.e. the approximation introduced in Sec. 4. As pointed out, this approximation corresponds to a purely hadronic description where all quark effects are suppressed. For instance, taking the static limit of the diagrams shown in Fig. 5, we exactly recover the diagrams which contribute to the pion electromagnetic form factor in the hadronic model of ref. . On the other hand, we already estimated that quark effects should not be negligible, i.e. we expect the static limit not to be a good approximation to the exact NJL calculations. In fact, if we take the parameters of our best fit to the pion electromagnetic form factor ($`\mathrm{\Lambda }_M`$ = 600 MeV) and insert the corresponding values for $`m`$ and $`f_\pi ^{(0)}`$ (see Table 1) into Eq. (40) we find $`g_{\rho \pi \pi }`$ = 9.3, which is considerably larger than the value of $`6`$, typically needed in hadronic models to describe the data. Fig. 8 displays the imaginary part of the $`\rho `$-meson propagator for $`\mathrm{\Lambda }_M`$ = 600 MeV. The solid line indicates the result of an exact treatment of our model. Since the parameters have been fitted to the pion electromagnetic form factor, the maximum is close to the empirical $`\rho `$-meson mass of 770 MeV. The cusp at $`\sqrt{s}`$ = 892 MeV is again a $`q\overline{q}`$-threshold effect. The corresponding result in the static limit is displayed by the dashed line. Of course there is no $`q\overline{q}`$-threshold in this approximation. Obviously the peak is much broader and shifted to higher energies. For a better understanding of this behavior we also compare the real- and imaginary parts of the corresponding $`\rho `$-meson self-energies (Fig. 9). Obviously, the differences seen in Fig. 8 are mainly due to the real part which is much more attractive in the exact NJL calculation than in the static limit, while the imaginary parts are not very different up to $`\sqrt{s}`$ 800 MeV. Note that, in this region, the imaginary part exclusively results from the $`\rho \pi \pi `$ process, i.e. quarks do not show up as unphysical decay channels. The differences to the static limit, both in the real- and in the imaginary part, come about by the fact, that the quarks cause a non-trivial momentum dependence of the intermediate meson propagators and of the effective meson-meson vertices. For instance, the quark triangles do not exactly behave as point-like three-meson vertices. This might be viewed as “physical quark effects” in contrast to the unphysical $`q\overline{q}`$-decays at higher energies. However, the momentum dependence of any quark loop below the $`q\overline{q}`$-threshold is related to the imaginary part above the threshold via dispersion relations. In that sense any difference to the static limit could be interpreted as an unphysical effect and one might ask whether “physical quark effects” exist at all. Of course, this fundamental question is far beyond the scope of our paper and can probably not be answered without understanding the mechanism of quark confinement itself. On the other hand, at sufficiently low energies ($`<`$ 400 MeV) the exact self energy of the $`\rho `$ meson does indeed almost coincide with the static limit. Here one can nicely see, how a hadronic model, which incorporates chiral symmetry, emerges naturally from the underlying quark structure once the many-body theory is carried to a sufficient degree of sophistication. ## 6 Conclusions We have investigated meson properties within the Nambu–Jona-Lasinio model, including meson-loop corrections, which are generated via a systematic $`1/N_c`$-expansion of the self energy in next-to-leading order. We have shown that such a scheme is consistent with chiral symmetry, leading to massless pions in the chiral limit. For non-vanishing current quark masses the pion mass is consistent with the Gell-Mann Oakes Renner relation if one carefully expands both sides of the relation up to next-to-leading order in $`1/N_c`$. The relative importance of the $`1/N_c`$-corrections is controlled by a parameter $`\mathrm{\Lambda }_M`$, which cuts off the three-momenta of the meson loops. One of the main goals of the present article was to determine the value of $`\mathrm{\Lambda }_M`$, together with the other parameters of the model. To that end we have performed a fit of the quark condensate $`\overline{\psi }\psi `$ and the pion decay constant $`f_\pi `$, together with the pion electromagnetic form factor $`F_\pi `$ in the time-like region. The latter more or less amounts to fitting the mass and the width of the $`\rho `$-meson. Here the meson loops are absolutely crucial in order to include the dominant $`\rho \pi \pi `$-decay channel, while the leading-order approximation contains only unphysical $`q\overline{q}`$-decay channels. Of course, a priori it was not clear to what extent these unphysical decay modes, which are an unavoidable consequence of the missing confinement mechanism in the NJL model, are still present in the region of the $`\rho `$-meson peak. It turns out that a reasonable fit of the above observables can be achieved with $`\mathrm{\Lambda }_M`$ = 600 MeV. For the constituent quark mass we find $`m`$ = 446 MeV. Hence the unphysical $`q\overline{q}`$-decay channel opens at 892 MeV, about 120 MeV above the maximum of the $`\rho `$-meson peak. Our result is also interesting in the context of Ref. where we have reported on the existence of instabilities in the pion sector for very large values of $`\mathrm{\Lambda }_M`$. However, with $`\mathrm{\Lambda }_M`$ = 600 MeV, as determined by our parameter fit, we are far away from this region. In fact, we find only moderate changes in the pion and quark sector: $`f_\pi `$ and $`\overline{\psi }\psi `$ are lowered by about 20% by the meson loop corrections, while the pion mass is increased by about 10%. This also indicates that our scheme converges rapidly and higher-order terms in the $`1/N_c`$-expansion are small. We have discussed that a hadronic model can be derived from our model if one takes the so-called static limit. Basically, this corresponds to a low-momentum approximation to the quark loops. In this way one can study the importance of quark effects, which are present in the exact model because of lack of confinement, but absent in the static limit. At higher energies we find for the $`\rho `$-meson rather large differences between the exact propagator and the static limit, whereas at sufficiently low energies ($`<`$ 400 MeV) the exact calculation and the static limit almost coincide. This implies that in the low-energy region the unphysical quark effects are suppressed and demonstrates that a hadronic model with realistic parameters indeed emerges from the underlying quark structure. It is likely that the quark threshold and hence the unphysical effects from continuum $`\overline{q}q`$ decay can be pushed to even higher energies once intermediate $`\rho `$-and $`a_1`$ meson states are included. Given the successful vacuum description of the $`\rho `$ meson, as presented in the present paper, it will be interesting to extend the calculation to finite temperature to asses medium modifications in the presence of a thermal heat bath. Because of spontaneous chiral symmetry breaking and its restoration at high temperature this implies a simultaneous treatment of the $`a_1`$ meson. Work in this direction is in progress. ## Acknowledgements We are indebted to G.J. van Oldenborgh for his assistance in questions related to his program package FF (see http://www.xs4all.nl/$``$gjvo/FF.html), which was used in parts of our numerical calculations. We also thank M. Urban for illuminating discussions. This work was supported in part by the BMBF and NSF grant NSF-PHY98-00978. ## Appendix A Definition of elementary integrals It is possible to reduce the expressions for the quark loops to some elementary integrals , see App. B and C. In this section we give the definitions of these integrals. $`I_1={\displaystyle \frac{d^4k}{(2\pi )^4}\frac{1}{k^2m^2+i\epsilon }},`$ (48) $`I(p)={\displaystyle \frac{d^4k}{(2\pi )^4}\frac{1}{(k^2m^2+i\epsilon )((k+p)^2m^2+i\epsilon )}},`$ (49) $`K(p)={\displaystyle \frac{d^4k}{(2\pi )^4}\frac{1}{(k^2m^2+i\epsilon )^2((k+p)^2m^2+i\epsilon )}},`$ (50) $`M(p_1,p_2)={\displaystyle \frac{d^4k}{(2\pi )^4}\frac{1}{(k^2m^2+i\epsilon )(k_1^2m^2+i\epsilon )(k_2^2m^2+i\epsilon )}},`$ (51) $`L(p_1,p_2,p_3)={\displaystyle \frac{d^4k}{(2\pi )^4}\frac{1}{(k^2m^2+i\epsilon )(k_1^2m^2+i\epsilon )(k_2^2m^2+i\epsilon )(k_3^2m^2+i\epsilon )}},`$ (52) $`p_1^\mu M_1(p_1,p_2)+p_2^\mu M_1(p_2,p_1)={\displaystyle \frac{d^4k}{(2\pi )^4}\frac{k^\mu }{(k^2m^2+i\epsilon )(k_1^2m^2+i\epsilon )(k_2^2m^2+i\epsilon )}},`$ (53) with $`k_i=k+p_i`$. The function $`M_1(p_1,p_2)`$ can be expressed in terms of the other integrals: $$M_1(p_1,p_2)=\frac{p_1p_2I(p_1)p_2^2I(p_2)+(p_2^2p_1p_2)I(p_1p_2)+p_2^2(p_1^2p_1p_2)M(p_1,p_2)}{2((p_1p_2)^2p_1^2p_2^2)},$$ (54) All integrals in Eqs. (48) to (53), are understood to be regularized. In our model we use Pauli-Villars regularization with two regulators, i.e. we replace $$\frac{d^4k}{(2\pi )^4}f(k;m)\frac{d^4k}{(2\pi )^4}\underset{j=0}{\overset{2}{}}c_jf(k;\mu _j),$$ (55) with $$\mu _j^2=m^2+j\mathrm{\Lambda }_q^2;c_0=1,c_1=2,c_2=1.$$ (56) One then gets the following relatively simple analytic expressions for the integrals $`I_1`$, $`I(p)`$ and $`K(p)`$: $`I_1={\displaystyle \frac{i}{16\pi ^2}}{\displaystyle \underset{j}{}}c_j\mu _j^2\mathrm{ln}(\mu _j^2)`$ (57) $`I(p)={\displaystyle \frac{i}{16\pi ^2}}{\displaystyle \underset{j}{}}c_j\left(x_{j1}\mathrm{ln}(x_{j1})+x_{j2}\mathrm{ln}(x_{j2})+x_{j1}\mathrm{ln}(p^2x_{j1})+x_{j2}\mathrm{ln}(p^2x_{j2})\right)`$ (58) $`I(p=0)={\displaystyle \frac{i}{16\pi ^2}}{\displaystyle \underset{j}{}}c_j\mathrm{ln}(\mu _j^2)`$ (59) $`K(p)={\displaystyle \frac{i}{16\pi ^2}}{\displaystyle \underset{j}{}}c_j{\displaystyle \frac{1}{2p^2(x_{j1}x_{j2})}}\left(\mathrm{ln}(x_{j1})\mathrm{ln}(x_{j1})+\mathrm{ln}(x_{j2})+\mathrm{ln}(x_{j2})\right),`$ (60) with $$x_{j1,2}=\frac{1}{2}\pm \frac{1}{2}\sqrt{1\frac{4\mu _j^2}{p^2}}.$$ (61) An analytic expression for the three-point function (Eq. 51) can be found in Refs. and . In certain kinematical regions the four-point function (eq. 52) is also known analytically . ## Appendix B RPA propagators Using the definitions given in the previous section the gap equation (Eq. (2)) takes the form $$m=m_0+2ig_s4N_cN_fmI_1.$$ (62) Similarly one can evaluate the quark-antiquark polarization diagrams (Eq. (3)) and calculate the RPA meson propagators. The results read: $`D_\sigma (p)`$ $`=`$ $`{\displaystyle \frac{2g_s}{\frac{m_0}{m}+2ig_s2N_cN_f(p^24m^2)I(p)}},`$ (63) $`D_\pi (p)`$ $`=`$ $`{\displaystyle \frac{2g_s}{\frac{m_0}{m}+2ig_s2N_cN_fp^2I(p)}},`$ (64) $`D_\rho (p)`$ $`=`$ $`{\displaystyle \frac{2g_v}{1+2ig_v\frac{4}{3}N_cN_f(2m^2I(0)+(p^2+2m^2)I(p))}},`$ (65) $`D_{a_1}(p)`$ $`=`$ $`{\displaystyle \frac{2g_v}{1+2ig_v\frac{4}{3}N_cN_f(2m^2I(0)+(p^24m^2)I(p))}}.`$ (66) We should comment on the $`\rho `$\- and the $`a_1`$-propagator. A straight-forward evaluation of the vector polarization diagrams gives $$\mathrm{\Pi }_\rho (p)=i\frac{4}{3}N_cN_f(2I_1+(p^2+2m^2)I(p)).$$ (67) Because of vector current conservation this function should vanish for $`p^2`$ = 0. This is only true if $$m^2I(0)=I_1,$$ (68) which is not the case if we regularize $`I(p)`$ and $`I_1`$ as described in App. A. This corresponds to the standard form of Pauli-Villars regularization in the NJL model . Alternatively one could perform the replacement Eq. (55) for the entire polarization loop. In fact, this is more in the original sense of Pauli-Villars regularization . Then the factor $`m^2`$ in Eq. (67) should be replaced by a factor $`\mu _j^2`$ inside the sum over regulators and one can easily show that Eq. (68) holds (see Eqs. (57) and (59)). However, this scheme would lead to even more severe problems: From the gap equation (Eq. 62) we conclude that $`iI_1`$ should be positive. On the other hand the pion decay constant in the chiral limit and in leading order in $`1/N_c`$ is given by $$f_\pi ^{2(0)}=2iN_cN_fm^2I(0).$$ (69) which implies that $`im^2I(0)`$ should be negative. So irrespective of the regularization scheme Eq. (68) cannot be fulfilled if we want to get reasonable results for $`m`$ and $`f_\pi ^{(0)}`$ at the same time. Therefore we choose the standard form of Pauli-Villars regularization in the NJL-model and replace the term $`I_1`$ in Eq. (67) by hand by $`m^2I(0)`$. This leads to the $`\rho `$-meson propagator as given in Eq. (65). For consistency the $`a_1`$ has been treated in the analogous way. ## Appendix C Explicit expressions for the meson-meson vertices In this section we list the explicit formulae for the meson-meson vertices. We restrict ourselves to those combinations which are needed for the calculations presented in this article. We begin with the three-meson vertices $`\mathrm{\Gamma }_{M_1,M_2,M_3}(q,p)`$ (see Fig. 4): $`i\mathrm{\Gamma }_{\sigma ,\sigma ,\sigma }(q,p)`$ $`=`$ $`i2mN\left(I(p^{})+I(q)+I(p)+(4m^2{\displaystyle \frac{1}{2}}(p^2+p^2+q^2))M(p,q)\right),`$ $`i\mathrm{\Gamma }_{\pi ,\pi ,\sigma }^{ab}(q,p)`$ $`=`$ $`i2mN\delta _{ab}\left(I(p^{})+pqM(p,q)\right),`$ $`i\mathrm{\Gamma }_{\rho ,\rho ,\sigma }^{\mu \lambda ,ab}(q,p)`$ $`=`$ $`\delta _{ab}h(q,p)\left(g^{\mu \lambda }{\displaystyle \frac{p^2q^\mu q^\lambda +q^2p^\mu p^\lambda pq(p^\mu q^\lambda +q^\mu p^\lambda )}{p^2q^2(pq)^2}}\right),`$ $`h(q,p)`$ $`=`$ $`imN\left(I(q)+I(p)2I(p^{})+(4m^22pqp^2)M(p,q)\right),`$ $`i\mathrm{\Gamma }_{\pi ,\pi ,\rho }^{\mu ,abc}(q,p)`$ $`=`$ $`\epsilon _{abc}\left(q^\mu f(q,p)p^\mu f(p,q)\right),`$ $`f(q,p)`$ $`=`$ $`N\left(I(q)+p^2M(p,q)+2pqM_1(q,p)\right),`$ (70) with $`p^{}=pq`$ and $`N=4N_cN_f`$. For the four-meson vertices we only need to consider the special cases needed for the diagrams (b) and (c) in Fig. 3: $`i\mathrm{\Gamma }_{\sigma ,\sigma ,\sigma ,\sigma }(q,p,q)`$ $`=`$ $`N\{{\displaystyle \frac{I(pq)+I(p+q)}{2}}+4m^2(M(p,q)+M(p,q))`$ $`+2(m^2(4m^2p^2q^2){\displaystyle \frac{p^2q^2}{4}})L(p,q,pq)\}`$ $`i\mathrm{\Gamma }_{\sigma ,\sigma ,\sigma ,\sigma }(q,p,p)`$ $`=`$ $`N\{I(p+q)+I(0)+4m^2(K(p)+K(q)+2M(p,q))`$ $`+2pqM(p,q)q^2K(q)p^2K(p)`$ $`+m^2(16m^24p^24q^2+{\displaystyle \frac{p^2q^2}{m^2}})L(p,q,0)\}`$ $`i\mathrm{\Gamma }_{\sigma ,\pi ,\sigma ,\pi }^{ab}(q,p,q)`$ $`=`$ $`\delta _{ab}N\left\{I(p+q)+I(pq)+p^2(4m^2q^2)L(p,q,pq)\right\}`$ $`i\mathrm{\Gamma }_{\sigma ,\pi ,\pi ,\sigma }^{ab}(q,p,p)`$ $`=`$ $`\delta _{ab}N\{I(p+q)I(0)(4m^2q^2)(K(q)p^2L(p,q,0))`$ $`+p^2K(p)2pqM(p,q)\}`$ $`i\mathrm{\Gamma }_{\pi ,\pi ,\pi ,\pi }^{abcd}(q,p,q)`$ $`=`$ $`N\kappa _{abcd}\left\{I(p+q)+I(pq)p^2q^2L(p,q,pq)\right\}`$ $`i\mathrm{\Gamma }_{\pi ,\pi ,\pi ,\pi }^{abcd}(q,p,p)`$ $`=`$ $`N\kappa _{abcd}\{I(p+q)+I(0)p^2K(p)`$ $`q^2K(q)+2pqM(p,q)+p^2q^2L(p,q,0)\}`$ $`i\mathrm{\Gamma }_{\rho ,\sigma ,\rho ,\sigma }^{ab}(q,p,q)`$ $`=`$ $`2\delta _{ab}N\{I(p+q)+I(pq)+2I(q)pq(M(p,q)M(p,q))`$ $`+(4m^22p^2)(M(p,q)+M(p,q))`$ $`+m^2(8m^26p^2+4q^2+{\displaystyle \frac{p^4(pq)^2}{m^2}})L(p,q,pq)\}`$ $`i\mathrm{\Gamma }_{\rho ,\sigma ,\sigma ,\rho }^{ab}(q,p,q)`$ $`=`$ $`2\delta _{ab}N\{I(p+q)I(0)+(p^24m^2)K(p)`$ $`+(q^2+2m^2)K(q)+(4m^22pq)M(p,q)`$ $`+m^2(8m^22p^2+4q^2{\displaystyle \frac{p^2q^2}{m^2}})L(p,q,0)\}`$ $`i\mathrm{\Gamma }_{\rho ,\pi ,\rho ,\pi }^{abcd}(q,p,q)`$ $`=`$ $`2N\kappa _{abcd}\{I(p+q)I(pq)2I(q)`$ $`+2p^2(M(p,q)+M(p,q))+pq(M(p,q)M(p,q))`$ $`+(2m^2p^2p^4+(pq)^2)L(p,q,pq)\}`$ $`i\mathrm{\Gamma }_{\rho ,\pi ,\pi ,\rho }^{abcd}(q,p,q)`$ $`=`$ $`2N\kappa _{abcd}\{I(p+q)+I(0)p^2K(p)(q^2+2m^2)K(q)`$ (71) $`+2pqM(p,q)+p^2(2m^2+q^2)L(p,q,0)\},`$ with $`\mathrm{\Gamma }_{\rho ,M,M,\rho }(q,p,q)=g_{\mu \nu }\mathrm{\Gamma }_{\rho ,M,M,\rho }^{\mu \nu }(q,p,q),\mathrm{\Gamma }_{\rho ,M,\rho ,M}(q,p,p)=g_{\mu \nu }\mathrm{\Gamma }_{\rho ,M,\rho ,M}^{\mu \nu }(q,p,p)`$ and $`\kappa _{abcd}=\delta _{ab}\delta _{cd}+\delta _{ad}\delta _{bc}\delta _{ac}\delta _{bd}`$.
warning/0001/hep-ph0001100.html
ar5iv
text
# QED radiative corrections to virtual Compton scattering ## I Introduction Virtual Compton scattering (VCS) has become in recent years a new and versatile tool in the study of nucleon structure and has triggered an important activity on both the theoretical and experimental side (see e.g. ). VCS, which is accessed through the $`(e,e^{}\gamma )`$ reaction, is studied now in various kinematical domains. At low energy, below pion production threshold, it allows to access generalized polarizabilities of the proton . These response functions, which constitute new nucleon structure observables, have been calculated in various approaches and models . To extract this nucleon structure information from VCS below pion production threshold, a considerable experimental effort is taking place at various electron laboratories. The first few events of VCS were observed in . The first dedicated VCS experiment has been performed at MAMI and for the first time, two combinations of generalized polarizabilities have been determined at a four-momentum squared $`Q^2`$ = 0.33 GeV<sup>2</sup> . An experiment at higher $`Q^2`$ (1 - 2 GeV<sup>2</sup>) at JLab has already been performed, which is under analysis at the time of writing, and a further experiment at lower $`Q^2`$ is planned at MIT-Bates . The VCS is also studied vigorously in the Bjorken regime (where the photon virtuality $`Q^2`$ and the photon-proton c.m. energy $`\sqrt{s}`$ are both large, with $`Q^2/s`$ finite), which is referred to as deeply virtual Compton scattering (DVCS). In this region, the DVCS amplitude is parametrized at leading order in $`Q`$ in terms of six generalized parton distributions , commonly denoted as skewed parton distributions (4 quark helicity conserving functions and 2 which involve a quark helicity flip). This field has generated by now a whole theoretical industry, and first experiments of DVCS and related hard electroproduction reactions are being performed, analyzed or planned both at JLab , HERMES/HERA , and COMPASS . The first absolute measurement of the VCS cross section on the nucleon performed at MAMI , indicates that QED radiative corrections provide an important contribution to the $`epep\gamma `$ reaction (of the order of 20% in the kinematics considered in ). The $`epep\gamma `$ reaction is particular in comparison with other electron scattering reactions because the photon can be emitted from both the proton side (this is the VCS process which contains the nucleon structure information of interest) or can be emitted from one of the electrons (which is the parasitic Bethe-Heitler process). The radiative corrections to the Bethe-Heitler process are formally different compared with the case of electron scattering. The importance of a very good understanding of the radiative corrections is indispensable if one wants to extract nucleon structure information from the $`epep\gamma `$ reaction, especially in those kinematical situations where the Bethe-Heitler process is not negligible. The calculation of these QED radiative corrections to the $`epep\gamma `$ reaction to first order in $`\alpha _{em}e^2/4\pi 1/137.036`$, is the subject of this paper. Radiative corrections were first calculated by Schwinger for potential scattering . Tsai extented the calculation of the radiative corrections to electron-proton scattering. The field has a long history and we refer to the standard review papers , which were used in the interpretation of many electron scattering experiments. The outline of the present paper is as follows. In section II, we introduce the kinematics and notations used for the $`epep\gamma `$ reaction, and give the lowest order amplitudes. In section III, we give the first order QED radiative corrections to the $`epep\gamma `$ reaction. We first calculate, in section III A, the one-loop virtual radiative corrections originating from the lepton side, to the $`epep\gamma `$ reaction. Our strategy used to evaluate the rather complicated loop integrals, is to solve first simpler loop integrals, which contain entirely the ultraviolet (UV) and infrared (IR) divergences, and in which the lowest order amplitudes factorize. These simpler loop integrals are evaluated analytically. The finite remainder with respect to the original amplitude, is then expressed through Feynman parameter integrals, which are calculated numerically in this work. In section III B, we calculate the soft photon emission contributions from the lepton side, to the $`epep\gamma `$ reaction. We discuss the similarities and differences with the bremsstrahlung contribution to elastic electron-nucleon scattering. These bremsstrahlung processes contain IR divergences which are shown to cancel exactly the IR divergences from the virtual photon processes. In section III C, the numerical method to evaluate the remaining finite Feynman parameter integrals is presented. We discuss subsequently the cases where the integrand is regular or singular, the latter originating from the propagation of on-shell intermediate states in the one-loop corrections to the $`epep\gamma `$ reaction. In particular, we discuss the different numerical checks performed and the accuracy of the calculation. In section III D, we discuss the radiative corrections at the proton side and the two-photon exchange corrections, by referring to the recent work of Maximon and Tjon . In section IV, we give a full numerical calculation for elastic electron-proton scattering of the photon emission processes where the photon energy is not very small compared with the lepton momenta, and which makes up the radiative tail. We compare this full calculation with an approximate procedure based on the angular peaking approximation, and show to what extent the full calculation validates the approximate method for the case of elastic electron-nucleon scattering. The approximate method will be seen to be realistic enough to apply it next to the calculation of the radiative tail in the case of VCS. In section V, we start by briefly discussing the radiative corrections to elastic electron-proton scattering. We apply the radiative corrections to elastic scattering data on the proton. We next give our results for the $`epep\gamma `$ reaction, and indicate how the observables are modified due to the first order QED radiative corrections. We discuss first the polarizability region for the $`epep\gamma `$ reaction, corresponding with a low outgoing photon energy. We show results for both unpolarized and polarized cross sections in MAMI and JLab kinematics. Subsequently, we give the effect of the first order QED radiative corrections to the DVCS cross section and the electron single spin asymmetry. Finally, we give our conclusions in section VI. We present technical details needed in the calculations, in two appendices. In appendix A, we calculate the radiative corrections to elastic lepton-nucleon scattering, which serves as a point of comparison with the $`epep\gamma `$ reaction. In particular, we present the details of the calculation of the soft photon emission contributions, and perform analytically the phase space integral over the soft photon in an exact way. We compare with other calculations in the literature. In appendix B, we present some technical details on the integration method used to evaluate singular Feynman parameter integrals. ## II Lowest order amplitudes of the $`epep\gamma `$ reaction The lowest order (in $`\alpha _{em}`$), contributions to the $`epep\gamma `$ reaction are given by the one-photon exchange processes. We denote in this work the four-momenta of the initial and final electrons by $`k(E_e,\stackrel{}{k}_e)`$ and $`k^{}(E_e^{},\stackrel{}{k}_e^{^{}})`$; the four-momenta of the initial and final protons by $`p(E_N,\stackrel{}{p}_N)`$ and $`p^{}(E_N^{},\stackrel{}{p}_N^{^{}})`$; and the four-momentum of the outgoing photon by $`q^{}(|\stackrel{}{q}^{^{}}|,\stackrel{}{q}^{^{}})`$. Furthermore, we denote $`qkk^{}=p^{}p+q^{}`$ and $`Q^2=q^2>0`$. The masses of the electron and proton are denoted by $`m`$ and $`M_N`$ respectively. The helicities of the initial (final) electrons are denoted by $`h(h^{})`$; the spins of initial (final) protons by $`s_p(s_p^{})`$; and the polarization four-vector of the outgoing photon by $`\epsilon `$. The spinors of initial and final electrons are denoted by $`u(k,h)`$ and $`u(k^{},h^{})`$; whereas the spinors of initial and final protons are denoted by $`N(p,s_p)`$ and $`N(p^{},s_p^{})`$. Throughout this work, we follow the conventions of Bjorken and Drell . In Figs. 1 (BHi) and (BHf), which are known as the Bethe-Heitler (BH) diagrams, a photon is emitted by either the incident or final electrons. The expressions for Figs. 1 (BHi) and (BHf) are respectively given by : $`M_{BH}^i=`$ $`ie^3\overline{u}(k^{^{}},h^{})\gamma ^\nu {\displaystyle \frac{\left(\mathit{}\mathit{}^{^{}}+m\right)}{2k.q^{^{}}}}\mathit{\epsilon ̸}^{}u(k,h){\displaystyle \frac{1}{\left(p^{^{}}p\right)^2}}\overline{N}(p^{^{}},s_p^{^{}})\mathrm{\Gamma }_\nu (p^{^{}},p)N(p,s_p),`$ (1) $`M_{BH}^f=`$ $`ie^3\overline{u}(k^{^{}},h^{})\mathit{\epsilon ̸}^{}{\displaystyle \frac{\left(\mathit{}^{^{}}+\mathit{}^{^{}}+m\right)}{2k^{^{}}.q^{^{}}}}\gamma ^\nu u(k,h){\displaystyle \frac{1}{\left(p^{^{}}p\right)^2}}\overline{N}(p^{^{}},s_p^{^{}})\mathrm{\Gamma }_\nu (p^{^{}},p)N(p,s_p),`$ (2) where the electron charge is given by $`(e)`$ (i.e. $`e>0`$ in this work). The on-shell electromagnetic vertex at the hadron side $`\mathrm{\Gamma }_\nu `$ in Eqs. (1,2) is given by $$\mathrm{\Gamma }_\nu (p^{^{}},p)=F_1\left((p^{^{}}p)^2\right)\gamma _\nu +F_2\left((p^{^{}}p)^2\right)i\sigma _{\nu \lambda }\frac{(p^{^{}}p)^\lambda }{2M_N},$$ (3) where $`F_1`$ and $`F_2`$ are respectively the Dirac and Pauli electromagnetic (on-shell) form factors of the nucleon. The four-momentum squared of the virtual photon in the BH processes is $`t=(p^{}p)^2`$, in contrast to $`q^2`$, which is the four-momentum squared for the VCS process $`\gamma ^{}p\gamma p`$, where the final photon is emitted from the hadron side. This latter part contains the nucleon structure information. The amplitude of the VCS contribution to the $`e^{}pe^{}p\gamma `$ reaction is given by : $$M_{VCS}=ie^3\overline{u}(k^{^{}},h^{})\gamma _\nu u(k,h)\frac{1}{q^2}\epsilon _\mu ^{}H^{\mu \nu }.$$ (4) Remark that for a positive lepton, the VCS amplitude changes sign. In Eq. (4), the gauge-invariant, hadronic tensor $`H^{\mu \nu }`$ is defined by : $$H^{\mu \nu }=id^4x\mathrm{e}^{iq.x}<p^{}|T[j^\nu (x),j^\mu (0)]|p>,$$ (5) where $`T`$ represents the time ordering, and $`j^\nu `$ the electromagnetic current operator. For the DVCS process in the Bjorken limit, the hadronic tensor of Eq. (5) is parametrized in terms of six leading twist skewed parton distributions (see e.g. ). For the VCS process at low energy, as investigated experimentally in , an important contribution to the tensor of Eq.(5) originates from the nucleon pole contributions shown in Figs. 1 (BORNi) and (BORNf). The contributions of the Born diagrams to the hadronic tensor are given by : $`H_{BORN,i}^{\mu \nu }=`$ $`\overline{N}(p^{^{}},s_p^{^{}})\mathrm{\Gamma }^\nu (p^{^{}},pq^{^{}}){\displaystyle \frac{\left(\mathit{}\mathit{}^{^{}}+M_N\right)}{2p.q^{^{}}}}\mathrm{\Gamma }^\mu (pq^{^{}},p)N(p,s_p),`$ (6) $`H_{BORN,f}^{\mu \nu }=`$ $`\overline{N}(p^{^{}},s_p^{^{}})\mathrm{\Gamma }^\mu (p^{^{}},p^{^{}}+q^{^{}}){\displaystyle \frac{\left(\mathit{}^{^{}}+\mathit{}^{^{}}+M_N\right)}{2p^{^{}}.q^{^{}}}}\mathrm{\Gamma }^\nu (p^{^{}}+q^{^{}},p)N(p,s_p),`$ (7) where the vertex $`\mathrm{\Gamma }^\mu `$ is now evaluated for off mass-shell values of one of its arguments. In Ref. , the Born diagrams were evaluated by using the vertex of Eq. (3). Doing so, the Born diagrams are separately gauge invariant. All nucleon structure effects are then absorbed in a non-Born amplitude which is regular in $`q^{}`$ and for which the Low Energy Theorem (LET) tells that it starts at order $`q^{}`$. The nucleon structure effects to the VCS tensor (Eq. (5)) below pion threshold, are then parametrized at order $`q^{}`$ in terms of six generalized (i.e. $`Q^2`$ dependent) nucleon polarizabilities . ## III First order radiative corrections to the $`epep\gamma `$ reaction ### A Virtual radiative corrections In this section, we calculate the one-loop QED virtual radiative corrections to the $`epep\gamma `$ reaction, which are represented in Fig. 2. In the present section, we consider only the corrections originating from the electron side as they can be calculated model-independently. The corrections originating from the hadronic side, for which a nucleon structure model is needed, will be discussed and estimated in section III D. The virtual radiative corrections to the BH process contain vertex corrections : Figs. 2 (V1i - V3i) and (V1f - V3f); electron self-energy corrections : Figs. 2 (Si, Sf); and vacuum polarization corrections : Figs. 2 (P1i, P1f). We indicate in our notation of the different diagrams whether the photon in the $`epep\gamma `$ reaction is emitted from the initial (i) electron or from the final (f) electron. The part of the virtual radiative corrections to the VCS process (i.e. where the photon in the reaction $`epep\gamma `$ is emitted from the hadronic side) which can be calculated model-independently, consists of the vertex diagram shown in Fig. 2 (V4) and the vacuum polarization diagram shown in Fig. 2 (P2). The blob in those figures represents the VCS process. For VCS below pion threshold, the blob is given by the Born diagrams (Fig. 1 (BORNi) and (BORNf)) + non-Born diagrams, which describe the nucleon polarizability effects. For DVCS, the blob is given in leading order by the so-called handbag diagrams, where the photon hits a quark in the proton . The calculation of the virtual radiative corrections to the VCS process is similar to that for electron scattering. The virtual radiative corrections to the Bethe-Heitler process are different, but involve the same one-loop building blocks, i.e. electron vertex, electron self-energy and photon self-energy. Therefore, we give in appendix A (sections A 1 \- A 4) the derivation and the expressions for these basic building blocks, and we apply it to elastic electron-nucleon scattering. In our calculations, we use the dimensional regularization method to treat both ultraviolet (UV) and infrared (IR) divergences. This amounts to evaluate all loop integrals in $`D`$ dimensions. The divergences then show up (when one takes $`D4`$) as poles of the form $`1/\epsilon `$, where $`\epsilon 2D/2`$. UV divergences are regularized by taking $`D<4`$ (i.e. $`\epsilon _{UV}=2D/2>0`$), whereas IR divergences are regularized by taking $`D>4`$ (i.e. $`\epsilon _{IR}=2D/2<0`$). Care has to be taken as to isolate the UV and IR divergent parts in the loop integrals first, as two different limits are understood when one takes $`D=4`$ at the end. The technical details of our calculational method can also be found in appendix A. We apply it here to calculate the diagrams of Fig. 2 to the $`epep\gamma `$ reaction. #### 1 Vertex correction diagrams of Figs. 2 (V1i) and (V1f) The amplitude corresponding to Fig. 2 (V1i) is given by $`M_{V1}^i={\displaystyle \frac{e^5}{\left(p^{^{}}p\right)^2}}\overline{N}(p^{^{}},s_p^{^{}})\mathrm{\Gamma }_\nu (p^{^{}},p)N(p,s_p)`$ (8) $`\times \overline{u}(k^{^{}},h^{})\gamma ^\nu {\displaystyle \frac{\left(\mathit{}\mathit{}^{^{}}+m\right)}{2k.q^{^{}}}}\mu ^{4D}{\displaystyle \frac{d^Dl}{\left(2\pi \right)^D}\frac{\gamma ^\alpha \left(\mathit{}\mathit{}^{^{}}\mathit{}+m\right)\mathit{\epsilon ̸}^{}\left(\mathit{}\mathit{}+m\right)\gamma _\alpha }{\left[l^2\right][l^22l.k][l^22l.(kq^{^{}})2k.q^{^{}}]}u(k,h)},`$ (9) where a mass scale $`\mu `$ (renormalization scale) is introduced when passing to $`D4`$ dimensions in order to keep the coupling constant dimensionless. One sees by inspection that the loop integral in Eq. (8), when taking $`D=4`$, is IR finite ($`l0`$ behavior), but has an UV divergence ($`l\mathrm{}`$ behavior). Our strategy to evaluate a complicated loop integral as in Eq. (8), is to solve first a simpler loop integral which contains entirely the UV divergence and which can be done analytically more easily. We observe from Eq. (8) that only the term in the numerator proportional to $`\gamma ^\alpha \mathit{}\mathit{\epsilon ̸}^{}\mathit{}\gamma _\alpha `$ is responsible for the UV divergence. To evaluate it, we add a similar term by replacing one factor in the denominator and evaluate this term analytically. In order to obtain the equivalence with $`M_{v1}^i`$, we have to subtract the added term again from the expression of Eq. (8). This leads to $`M_{V1}^i={\displaystyle \frac{e^5}{\left(p^{^{}}p\right)^2}}\overline{N}(p^{^{}},s_p^{^{}})\mathrm{\Gamma }_\nu (p^{^{}},p)N(p,s_p)`$ (10) $`\times \overline{u}(k^{^{}},h^{})\gamma ^\nu {\displaystyle \frac{\left(\mathit{}\mathit{}^{^{}}+m\right)}{2k.q^{^{}}}}\{\mu ^{4D}{\displaystyle }{\displaystyle \frac{d^Dl}{\left(2\pi \right)^D}}{\displaystyle \frac{\gamma ^\alpha \mathit{}\mathit{\epsilon ̸}^{}\mathit{}\gamma _\alpha }{\left[l^2\right][l^22l.k][l^22l.k^{^{}}]}}`$ (11) $`+2{\displaystyle \frac{d^4l}{\left(2\pi \right)^4}\frac{\mathit{}\mathit{\epsilon ̸}^{}\left(\mathit{}\mathit{}^{^{}}\right)+\mathit{}\mathit{\epsilon ̸}^{}\mathit{}m^2\mathit{\epsilon ̸}^{}+4m\epsilon ^{}.\left(kl\right)\mathit{}\mathit{\epsilon ̸}^{}\left(\mathit{}\mathit{}^{^{}}\right)}{\left[l^2\right][l^22l.k][l^22l.(kq^{^{}})2k.q^{^{}}]}}`$ (12) $`+2{\displaystyle }{\displaystyle \frac{d^4l}{\left(2\pi \right)^4}}{\displaystyle \frac{2\mathit{}\mathit{\epsilon ̸}^{}\mathit{}[l.(qq^{^{}})+k.q^{^{}}]}{\left[l^2\right][l^22l.k][l^22l.k^{^{}}][l^22l.(kq^{^{}})2k.q^{^{}}]}}\}u(k,h).`$ (13) It should be remarked that only the added term in Eq. (10) (first term within curly brackets of Eq. (10) ) is UV divergent and has therefore to be evaluated in $`D`$ dimensions using the dimensional regularization method. The third term within the curly brackets in Eq. (10) is the difference between the term proportional to $`\gamma ^\alpha \mathit{}\mathit{\epsilon ̸}^{}\mathit{}\gamma _\alpha `$ in Eq. (8) and the added term (which has one different factor in the denominator). As can be seen by power counting, this term is UV finite and can therefore readily be evaluated for $`D=4`$. The denominator in the UV divergent first term of Eq. (10), was chosen so that it corresponds with the vertex correction which appears in electron scattering. Therefore, this UV divergent term can be calculated analytically along a similar way as was performed in appendix A. The result is given by : $`\mu ^{4D}{\displaystyle \frac{d^Dl}{\left(2\pi \right)^D}\frac{\gamma ^\alpha \mathit{}\mathit{\epsilon ̸}^{}\mathit{}\gamma _\alpha }{\left[l^2\right][l^22l.k][l^22l.k^{^{}}]}}`$ (16) $`={\displaystyle \frac{i}{\left(4\pi \right)^2}}\{\mathit{\epsilon ̸}^{}[{\displaystyle \frac{1}{\epsilon _{UV}}}\gamma _E+\mathrm{ln}\left({\displaystyle \frac{4\pi \mu ^2}{m^2}}\right)+1v\mathrm{ln}\left({\displaystyle \frac{v+1}{v1}}\right)]{\displaystyle \frac{1}{Q^2}}\mathit{}\mathit{\epsilon ̸}^{}\mathit{}`$ $`+{\displaystyle \frac{1}{Q^2v}}\mathrm{ln}\left({\displaystyle \frac{v+1}{v1}}\right)[\mathit{}\mathit{\epsilon ̸}^{}\mathit{}^{^{}}+\mathit{}^{^{}}\mathit{\epsilon ̸}^{}\mathit{}+\left({\displaystyle \frac{v^2+1}{2}}\right)\mathit{}\mathit{\epsilon ̸}^{}\mathit{}]\},`$ where $`v`$ is defined as $$v^2\mathrm{\hspace{0.33em}1}+\frac{4m^2}{Q^2}.$$ (17) The UV divergence in Eq. (16) is removed by the corresponding vertex counterterm as given by Eqs. (A7,A19) $$(CT)_{V1}^i=M_{BH}^i\frac{(e^2)}{\left(4\pi \right)^2}\left\{\left[\frac{1}{\epsilon _{UV}}\gamma _E+\mathrm{ln}\left(\frac{4\pi \mu ^2}{m^2}\right)\right]+2\left[\frac{1}{\epsilon _{IR}}\gamma _E+\mathrm{ln}\left(\frac{4\pi \mu ^2}{m^2}\right)\right]+4\right\},$$ (18) where we have used the expression of Eq. (1) for the BH amplitude $`M_{BH}^i`$. Adding the counterterm of Eq. (18) to Eq. (10) and introducing a Feynman parametrization in the second and third terms of Eq. (10) in order to perform the integrals over $`l`$, yields the total, UV finite result : $`M_{V1}^i+(CT)_{V1}^i=M_{BH}^i{\displaystyle \frac{e^2}{\left(4\pi \right)^2}}\left\{2\left[{\displaystyle \frac{1}{\epsilon _{IR}}}\gamma _E+\mathrm{ln}\left({\displaystyle \frac{4\pi \mu ^2}{m^2}}\right)\right]3v\mathrm{ln}\left({\displaystyle \frac{v+1}{v1}}\right)\right\}`$ (19) $`+{\displaystyle \frac{ie^5}{\left(4\pi \right)^2}}{\displaystyle \frac{1}{\left(p^{^{}}p\right)^2}}\overline{N}(p^{^{}},s_p^{^{}})\mathrm{\Gamma }_\nu (p^{^{}},p)N(p,s_p)`$ (20) $`\times \overline{u}(k^{^{}},h^{})\gamma ^\nu {\displaystyle \frac{\left(\mathit{}\mathit{}^{^{}}+m\right)}{2k.q^{^{}}}}`$ (21) $`\times \{{\displaystyle \frac{1}{Q^2}}[(1+{\displaystyle \frac{v^2+1}{2v}}\mathrm{ln}\left({\displaystyle \frac{v+1}{v1}}\right))\mathit{}\mathit{\epsilon ̸}^{}\mathit{}+{\displaystyle \frac{1}{v}}\mathrm{ln}\left({\displaystyle \frac{v+1}{v1}}\right)\{\mathit{}\mathit{\epsilon ̸}^{}\mathit{}^{^{}}+\mathit{}^{^{}}\mathit{\epsilon ̸}^{}\mathit{}\}]`$ (22) $`\mathrm{\hspace{0.17em}2}{\displaystyle \underset{0}{\overset{1}{}}}dy{\displaystyle \underset{0}{\overset{1}{}}}dx{\displaystyle \frac{1}{B_1^i}}[y(\mathit{}\mathit{}^{^{}}x)\mathit{\epsilon ̸}^{}(\mathit{}\mathit{}^{^{}})+y\mathit{}\mathit{\epsilon ̸}^{}(\mathit{}\mathit{}^{^{}}x)`$ (23) $`+\mathrm{\hspace{0.17em}4}m(\epsilon ^{}.k)(1y)\mathit{}\mathit{\epsilon ̸}^{}(\mathit{}\mathit{}^{^{}})m^2\mathit{\epsilon ̸}^{}]`$ (24) $`\mathrm{\hspace{0.17em}4}{\displaystyle \underset{0}{\overset{1}{}}}dx_3x_3^2{\displaystyle \underset{0}{\overset{1}{}}}dx_2x_2{\displaystyle \underset{0}{\overset{1}{}}}dx_1[({\displaystyle \frac{1}{A^i}}\mathit{\epsilon ̸}^{}+{\displaystyle \frac{1}{\left(A^i\right)^2}}\mathit{}^i\mathit{\epsilon ̸}^{}\mathit{}^i)(P^i.(qq^{^{}})+k.q^{^{}})`$ (25) $`{\displaystyle \frac{1}{2A^i}}((\mathit{}\mathit{}^{^{}})\mathit{\epsilon ̸}^{}\mathit{}^i+\mathit{}^i\mathit{\epsilon ̸}^{}(\mathit{}\mathit{}^{^{}}))]\}u(k,h),`$ (26) with the four-vector $`P^i`$ defined by $$P^i\left(kq^{^{}}\right)(1x_3)+\left(kqx_1\right)x_2x_3,$$ (27) and the scalars $`A^i`$ and $`B^i`$ defined by $$A^i\mathrm{\hspace{0.33em}2}k.q^{^{}}(1x_3)+\left(P^i\right)^2,$$ (28) and $$B_1^i\mathrm{\hspace{0.33em}2}k.q^{^{}}x(1y)+m^2y.$$ (29) Remark that although Eq. (19) is UV finite, it contains now an IR divergence through the vertex counterterm of Eq. (18) as shown in appendix A (Eq. (A19)). We will demonstrate however in section III B, that all IR divergences, arising from the one-loop corrections to the $`epep\gamma `$ reaction, are cancelled when adding the corresponding soft photon emission contributions. The Feynman parameter integrals in Eq. (19) which orginate from the finite integrals in Eq. (10) remain to be evaluated. As an analytical calculation of these integrals is rather complicated, we will evaluate them numerically in this paper, which will be discussed in section III C. In a completely similar way as for Fig. 2 (V1i), the total amplitude including the counterterm corresponding to Fig. 2 (V1f) yields : $`M_{V1}^f+(CT)_{V1}^f=M_{BH}^f{\displaystyle \frac{e^2}{\left(4\pi \right)^2}}\left\{2\left[{\displaystyle \frac{1}{\epsilon _{IR}}}\gamma _E+\mathrm{ln}\left({\displaystyle \frac{4\pi \mu ^2}{m^2}}\right)\right]3v\mathrm{ln}\left({\displaystyle \frac{v+1}{v1}}\right)\right\}`$ (30) $`+{\displaystyle \frac{ie^5}{\left(4\pi \right)^2}}{\displaystyle \frac{1}{\left(p^{^{}}p\right)^2}}\overline{N}(p^{^{}},s_p^{^{}})\mathrm{\Gamma }_\nu (p^{^{}},p)N(p,s_p)`$ (31) $`\times \overline{u}(k^{},h^{})\{{\displaystyle \frac{1}{Q^2}}[(1+{\displaystyle \frac{v^2+1}{2v}}\mathrm{ln}\left({\displaystyle \frac{v+1}{v1}}\right))\mathit{}\mathit{\epsilon ̸}^{}\mathit{}+{\displaystyle \frac{1}{v}}\mathrm{ln}\left({\displaystyle \frac{v+1}{v1}}\right)\{\mathit{}\mathit{\epsilon ̸}^{}\mathit{}^{^{}}+\mathit{}^{^{}}\mathit{\epsilon ̸}^{}\mathit{}\}]`$ (32) $`\mathrm{\hspace{0.17em}2}{\displaystyle \underset{0}{\overset{1}{}}}dy{\displaystyle \underset{0}{\overset{1}{}}}dx{\displaystyle \frac{1}{B_1^f}}[y(\mathit{}^{^{}}+\mathit{}^{^{}})\mathit{\epsilon ̸}^{}(\mathit{}^{^{}}+\mathit{}^{^{}}x)+y(\mathit{}^{^{}}+\mathit{}^{^{}}x)\mathit{\epsilon ̸}^{}\mathit{}^{^{}}`$ (33) $`+\mathrm{\hspace{0.17em}4}m(\epsilon ^{}.k^{^{}})(1y)(\mathit{}^{^{}}+\mathit{}^{^{}})\mathit{\epsilon ̸}^{}\mathit{}^{^{}}m^2\mathit{\epsilon ̸}^{}]`$ (34) $`+\mathrm{\hspace{0.17em}4}{\displaystyle \underset{0}{\overset{1}{}}}dx_3x_3^2{\displaystyle \underset{0}{\overset{1}{}}}dx_2x_2{\displaystyle \underset{0}{\overset{1}{}}}dx_1[({\displaystyle \frac{1}{A^f}}\mathit{\epsilon ̸}^{}+{\displaystyle \frac{1}{\left(A^f\right)^2}}\mathit{}^f\mathit{\epsilon ̸}^{}\mathit{}^f)(P^f.(qq^{^{}})+k^{^{}}.q^{^{}})`$ (35) $`{\displaystyle \frac{1}{2A^f}}((\mathit{}\mathit{}^{^{}})\mathit{\epsilon ̸}^{}\mathit{}^f+\mathit{}^f\mathit{\epsilon ̸}^{}(\mathit{}\mathit{}^{^{}}))]`$ (36) $`{\displaystyle \frac{}{}}\}{\displaystyle \frac{\left(\mathit{}^{^{}}+\mathit{}^{^{}}+m\right)}{2k^{^{}}.q^{^{}}}}\gamma ^\nu u(k,h),`$ (37) with the four-vector $`P^f`$ defined by $$P^f\left(k^{^{}}+q^{^{}}\right)(1x_3)+\left(k^{^{}}+qx_1\right)x_2x_3,$$ (38) and the scalars $`A^f`$ and $`B_1^f`$ defined by $$A^f2k^{^{}}.q^{^{}}(1x_3)+\left(P^f\right)^2,$$ (39) and $$B_1^f2k^{^{}}.q^{^{}}x(1y)+m^2y.$$ (40) #### 2 Vertex correction diagrams of Figs. 2 (V2i) and (V2f) The amplitude corresponding to Fig. 2 (V2i) is given by $`M_{V2}^i={\displaystyle \frac{e^5}{\left(p^{^{}}p\right)^2}}\overline{N}(p^{^{}},s_p^{^{}})\mathrm{\Gamma }_\nu (p^{^{}},p)N(p,s_p)`$ (41) $`\times \overline{u}(k^{^{}},h^{})\mu ^{4D}{\displaystyle \frac{d^Dl}{\left(2\pi \right)^D}\frac{\gamma ^\alpha \left(\mathit{}^{}\mathit{}+m\right)\gamma ^\nu \left(\mathit{}\mathit{}^{^{}}\mathit{}+m\right)\gamma _\alpha }{\left[l^2\right][l^22l.k^{}][l^22l.(kq^{^{}})2k.q^{^{}}]}\frac{\left(\mathit{}\mathit{}^{^{}}+m\right)}{2k.q^{^{}}}\mathit{\epsilon ̸}^{}u(k,h)}.`$ (42) One sees from Eq. (41) that again only the term in the numerator proportional to $`\gamma ^\alpha \mathit{}\gamma ^\nu \mathit{}\gamma _\alpha `$ contains an UV divergence for $`D=4`$. To evaluate the loop integral of Eq. (41), we therefore apply a similar trick as used before in Eq. (10). This amounts to adding and subtracting a term in Eq. (41) by replacing $`(l^22l.\left(kq^{}\right)2k.q^{})`$ in the denominator by $`(l^22l.k^{})`$, and which contains entirely the UV divergence. The further steps are then analogous to those following Eq. (10), and yield the following result for Fig. 2 (V2i) : $`M_{V2}^i+(CT)_{V2}^i=M_{BH}^i{\displaystyle \frac{e^2}{\left(4\pi \right)^2}}\left\{2\left[{\displaystyle \frac{1}{\epsilon _{IR}}}\gamma _E+\mathrm{ln}\left({\displaystyle \frac{4\pi \mu ^2}{m^2}}\right)\right]3v\mathrm{ln}\left({\displaystyle \frac{v+1}{v1}}\right)\right\}`$ (43) $`+{\displaystyle \frac{ie^5}{\left(4\pi \right)^2}}{\displaystyle \frac{1}{\left(p^{^{}}p\right)^2}}\overline{N}(p^{^{}},s_p^{^{}})\mathrm{\Gamma }_\nu (p^{^{}},p)N(p,s_p)`$ (44) $`\times \overline{u}(k^{^{}},h^{})\{{\displaystyle \frac{1}{Q^2}}[(1+{\displaystyle \frac{v^2+1}{2v}}\mathrm{ln}\left({\displaystyle \frac{v+1}{v1}}\right))\mathit{}\gamma ^\nu \mathit{}+{\displaystyle \frac{1}{v}}\mathrm{ln}\left({\displaystyle \frac{v+1}{v1}}\right)\{\mathit{}\gamma ^\nu \mathit{}^{^{}}+\mathit{}^{^{}}\gamma ^\nu \mathit{}\}\}`$ (45) $`\mathrm{\hspace{0.17em}2}{\displaystyle \underset{0}{\overset{1}{}}}dy{\displaystyle \underset{0}{\overset{1}{}}}dx{\displaystyle \frac{1}{B_2^i}}[y(\mathit{}\mathit{}^{^{}})\gamma ^\nu (\mathit{}^{^{}}+(\mathit{}\mathit{}^{^{}})x)+y(\mathit{}^{^{}}+(\mathit{}\mathit{}^{^{}})x)\gamma ^\nu \mathit{}^{^{}}`$ (46) $`+\mathrm{\hspace{0.17em}4}m\left(k^{^{}}\right)^\nu (1y)(\mathit{}\mathit{}^{^{}})\gamma ^\nu \mathit{}^{^{}}m^2\gamma ^\nu ]`$ (47) $`\mathrm{\hspace{0.17em}4}{\displaystyle \underset{0}{\overset{1}{}}}dx_3x_3^2{\displaystyle \underset{0}{\overset{1}{}}}dx_2x_2{\displaystyle \underset{0}{\overset{1}{}}}dx_1[({\displaystyle \frac{1}{A^i}}\gamma ^\nu +{\displaystyle \frac{1}{\left(A^i\right)^2}}\mathit{}^i\gamma ^\nu \mathit{}^i)(q^{^{}}.(kP^i))`$ (48) $`+{\displaystyle \frac{1}{2A^i}}(\mathit{}^{^{}}\gamma ^\nu \mathit{}^i+\mathit{}^i\gamma ^\nu \mathit{}^{^{}})]`$ (49) $`{\displaystyle \frac{}{}}\}{\displaystyle \frac{\left(\mathit{}\mathit{}^{^{}}+m\right)}{2k.q^{^{}}}}\mathit{ϵ̸}^{}u(k,h),`$ (50) where $`A^i`$ is given as in Eq. (28) and where $$B_2^im^2y+x^2y\left(qq^{^{}}\right)^2+\mathrm{\hspace{0.17em}2}xk.q^{^{}}+2xyk^{^{}}.\left(qq^{^{}}\right).$$ (51) In an analogous way, the amplitude corresponding to Fig. 2 (V2f) can be calculated, and yields as result : $`M_{V2}^f+(CT)_{V2}^f=M_{BH}^f{\displaystyle \frac{e^2}{\left(4\pi \right)^2}}\left\{2\left[{\displaystyle \frac{1}{\epsilon _{IR}}}\gamma _E+\mathrm{ln}\left({\displaystyle \frac{4\pi \mu ^2}{m^2}}\right)\right]3v\mathrm{ln}\left({\displaystyle \frac{v+1}{v1}}\right)\right\}`$ (52) $`+{\displaystyle \frac{ie^5}{\left(4\pi \right)^2}}{\displaystyle \frac{1}{\left(p^{^{}}p\right)^2}}\overline{N}(p^{^{}},s_p^{^{}})\mathrm{\Gamma }_\nu (p^{^{}},p)N(p,s_p)`$ (53) $`\times \overline{u}(k^{^{}},h^{})\mathit{ϵ̸}^{}{\displaystyle \frac{\left(\mathit{}^{^{}}+\mathit{}^{^{}}+m\right)}{2k^{^{}}.q^{^{}}}}`$ (54) $`\times \{{\displaystyle \frac{1}{Q^2}}[(1+{\displaystyle \frac{v^2+1}{2v}}\mathrm{ln}\left({\displaystyle \frac{v+1}{v1}}\right))\mathit{}\gamma ^\nu \mathit{}+{\displaystyle \frac{1}{v}}\mathrm{ln}\left({\displaystyle \frac{v+1}{v1}}\right)\{\mathit{}\gamma ^\nu \mathit{}^{^{}}+\mathit{}^{^{}}\gamma ^\nu \mathit{}\}\}`$ (55) $`\mathrm{\hspace{0.17em}2}{\displaystyle \underset{0}{\overset{1}{}}}dy{\displaystyle \underset{0}{\overset{1}{}}}dx{\displaystyle \frac{1}{B_2^i}}[y\mathit{}\gamma ^\nu (\mathit{}(\mathit{}\mathit{}^{^{}})x)+y(\mathit{}(\mathit{}\mathit{}^{^{}})x)\gamma ^\nu (\mathit{}^{^{}}+\mathit{}^{^{}})`$ (56) $`+\mathrm{\hspace{0.17em}4}m\left(k\right)^\nu (1y)\mathit{}\gamma ^\nu (\mathit{}^{^{}}+\mathit{}^{^{}})m^2\gamma ^\nu ]`$ (57) $`+\mathrm{\hspace{0.17em}4}{\displaystyle \underset{0}{\overset{1}{}}}dx_3x_3^2{\displaystyle \underset{0}{\overset{1}{}}}dx_2x_2{\displaystyle \underset{0}{\overset{1}{}}}dx_1[({\displaystyle \frac{1}{A^f}}\gamma ^\nu +{\displaystyle \frac{1}{\left(A^f\right)^2}}\mathit{}^f\gamma ^\nu \mathit{}^f)(q^{^{}}.(k^{^{}}P^f))`$ (58) $`+{\displaystyle \frac{1}{2A^f}}(\mathit{}^{^{}}\gamma ^\nu \mathit{}^f+\mathit{}^f\gamma ^\nu \mathit{}^{^{}})]\}u(k,h),`$ (59) where $`A^f`$ is given as in Eq. (39) and where $$B_2^fm^2y+x^2y\left(qq^{^{}}\right)^2\mathrm{\hspace{0.17em}2}xk^{^{}}.q^{^{}}2xyk.\left(qq^{^{}}\right).$$ (60) #### 3 Vertex correction diagrams of Figs. 2 (V3i) and (V3f) The amplitude $`M_{V3}^i`$ corresponding to Fig. 2 (V3i) is given by $`M_{V3}^i={\displaystyle \frac{e^5}{\left(p^{^{}}p\right)^2}}\overline{N}(p^{^{}},s_p^{^{}})\mathrm{\Gamma }_\nu (p^{^{}},p)N(p,s_p)`$ (61) $`\times \overline{u}(k^{^{}},h^{})\mu ^{4D}{\displaystyle \frac{d^Dl}{\left(2\pi \right)^D}\frac{\gamma ^\alpha \left(\mathit{}^{^{}}+\mathit{}+m\right)\gamma ^\nu \left(\mathit{}\mathit{}^{^{}}+\mathit{}+m\right)\mathit{\epsilon ̸}^{}\left(\mathit{}+\mathit{}+m\right)\gamma _\alpha }{\left[l^2\right][l^2+2l.k^{^{}}][l^2+2l.k][l^2+2l.(kq^{^{}})2k.q^{^{}}]}u(k,h)}.`$ (62) Remark that the loop integral in Eq. (61) is UV finite but contains an IR divergence for $`D=4`$. This is because in Fig. 2 (V3i), a soft virtual photon ($`l0`$) couples to two on-shell electron lines. To isolate the IR divergence, we first decompose the numerator in Eq. (61) by using the relations $`\overline{u}(k^{^{}},h^{})\gamma ^\alpha \left(\mathit{}^{^{}}+m\right)=\overline{u}(k^{^{}},h^{})\mathrm{\hspace{0.17em}2}k_\alpha ^{^{}}`$ and $`\left(\mathit{}+m\right)\gamma _\alpha u(k,h)=\mathrm{\hspace{0.17em}2}k_\alpha u(k,h)`$. This yields : $`M_{V3}^i={\displaystyle \frac{e^5}{\left(p^{^{}}p\right)^2}}\overline{N}(p^{^{}},s_p^{^{}})\mathrm{\Gamma }_\nu (p^{^{}},p)N(p,s_p)`$ (66) $`\times \overline{u}(k^{^{}},h^{})\mu ^{4D}{\displaystyle \frac{d^Dl}{\left(2\pi \right)^D}\frac{1}{\left[l^2\right][l^2+2l.k^{^{}}][l^2+2l.k][l^2+2l.(kq^{^{}})2k.q^{^{}}]}}`$ $`\times \{4(k.k^{})\gamma ^\nu (\mathit{}\mathit{}^{^{}}+m)\mathit{\epsilon ̸}^{}+4(k.k^{})\gamma ^\nu \mathit{}\mathit{\epsilon ̸}^{}+2\gamma ^\nu (\mathit{}\mathit{}^{^{}}+\mathit{}+m)\mathit{\epsilon ̸}^{}\mathit{}\mathit{}^{^{}}`$ $`+2\mathit{}\mathit{}\gamma ^\nu (\mathit{}\mathit{}^{^{}}+\mathit{}+m)\mathit{\epsilon ̸}^{}+\gamma ^\alpha \mathit{}\gamma ^\nu (\mathit{}\mathit{}^{^{}}+\mathit{}+m)\mathit{\epsilon ̸}^{}\mathit{}\gamma _\alpha \}u(k,h).`$ In Eq. (66), only the term in the numerator which is $`l`$-independent (the first term within the curly brackets) contains an IR divergence, whereas all the other terms are finite. As before, instead of aiming at an analytical formula for a rather complicated integral, we evaluate the IR divergent part of the integral in Eq. (66) by adding and subtracting a term that contains the divergence and that can be performed analytically rather easily. In constructing this term, we are looking for a denominator which contains the same dependence as the basic BH process in order that this BH amplitude can be factored from this IR divergent term. This yields the following expression, which is by construction identical to Eq. (66) : $`M_{V3}^i={\displaystyle \frac{e^5}{\left(p^{^{}}p\right)^2}}\overline{N}(p^{^{}},s_p^{^{}})\mathrm{\Gamma }_\nu (p^{^{}},p)N(p,s_p)`$ (67) $`\times \overline{u}(k^{^{}},h^{})\{\mu ^{4D}{\displaystyle }{\displaystyle \frac{d^Dl}{\left(2\pi \right)^D}}{\displaystyle \frac{4(k.k^{})\gamma ^\nu (\mathit{}\mathit{}^{^{}}+m)\mathit{\epsilon ̸}^{}}{\left[l^2\right][l^2+2l.k^{^{}}][l^2+2l.k][2k.q^{^{}}]}}`$ (68) $`+{\displaystyle \frac{d^4l}{\left(2\pi \right)^4}\frac{1}{\left[l^2\right][l^2+2l.k^{^{}}][l^2+2l.k][l^2+2l.(kq^{^{}})2k.q^{^{}}]}}`$ (69) $`\times [4(k.k^{})\gamma ^\nu (\mathit{}\mathit{}^{^{}}+m)\mathit{\epsilon ̸}^{}{\displaystyle \frac{l^22l.\left(kq^{^{}}\right)}{2k.q^{^{}}}}+\mathrm{\hspace{0.17em}4}(k.k^{})\gamma ^\nu \mathit{}\mathit{\epsilon ̸}^{}`$ (70) $`+\mathrm{\hspace{0.17em}2}\gamma ^\nu \left(\mathit{}\mathit{}^{^{}}+\mathit{}+m\right)\mathit{\epsilon ̸}^{}\mathit{}\mathit{}^{^{}}+\mathrm{\hspace{0.17em}2}\mathit{}\mathit{}\gamma ^\nu \left(\mathit{}\mathit{}^{^{}}+\mathit{}+m\right)\mathit{\epsilon ̸}^{}`$ (71) $`+\gamma ^\alpha \mathit{}\gamma ^\nu (\mathit{}\mathit{}^{^{}}+\mathit{}+m)\mathit{\epsilon ̸}^{}\mathit{}\gamma _\alpha ]\}u(k,h).`$ (72) Remark that the added term (first term of Eq. (67)) contains the IR divergence whereas the other terms of Eq. (67)) do not have any divergences so that the corresponding integrals may be performed directly in four dimensions as indicated. For the first term of Eq. (67) we furthermore see that the $`l`$-independent part of the energy denominator is the same as the one occuring in the corresponding Bethe-Heitler diagram (Fig. 1a). The $`l`$-dependent part of the energy denominator for this term is the same as the one for the vertex correction to elastic electron scattering, Eq. (A10). The corresponding integral may therefore be evaluated analytically in a similar way as was done in appendix A. This yields for the IR divergent term in Eq. (67) : $`{\displaystyle \frac{e^5}{\left(p^{^{}}p\right)^2}}\overline{N}(p^{^{}},s_p^{^{}})\mathrm{\Gamma }_\nu (p^{^{}},p)N(p,s_p)`$ (73) $`\times \overline{u}(k^{^{}},h^{})\gamma ^\nu {\displaystyle \frac{\left(\mathit{}\mathit{}^{^{}}+m\right)}{2k.q^{^{}}}}\mathit{\epsilon ̸}^{}u(k,h)\mu ^{4D}{\displaystyle \frac{d^Dl}{\left(2\pi \right)^D}\frac{4(k.k^{^{}})}{\left[l^2\right][l^2+2l.k^{^{}}][l^2+2l.k]}}`$ (74) $`=M_{BH}^i{\displaystyle \frac{e^2}{\left(4\pi \right)^2}}\{[{\displaystyle \frac{1}{\epsilon _{IR}}}\gamma _E+\mathrm{ln}\left({\displaystyle \frac{4\pi \mu ^2}{m^2}}\right)]{\displaystyle \frac{v^2+1}{v}}\mathrm{ln}\left({\displaystyle \frac{v+1}{v1}}\right)`$ (75) $`+{\displaystyle \frac{v^2+1}{2v}}\mathrm{ln}\left({\displaystyle \frac{v+1}{v1}}\right)\mathrm{ln}\left({\displaystyle \frac{v^21}{4v^2}}\right)+{\displaystyle \frac{v^2+1}{v}}[Sp\left({\displaystyle \frac{v+1}{2v}}\right)Sp\left({\displaystyle \frac{v1}{2v}}\right)]\}.`$ (76) The evaluation of the finite four-dimensional integral in Eq. (67) can be performed at the expense of the introduction of three Feynman parameter integrals due to the four energy denominators : $`{\displaystyle \frac{1}{\left[l^2\right][l^2+2l.k^{^{}}][l^2+2l.k][l^2+2l.(kq^{^{}})2k.q^{^{}}]}}`$ (77) $`=\mathrm{\hspace{0.17em}6}{\displaystyle \underset{0}{\overset{1}{}}}𝑑yy^2{\displaystyle \underset{0}{\overset{1}{}}}𝑑x_2x_2{\displaystyle \underset{0}{\overset{1}{}}}𝑑x_1{\displaystyle \frac{1}{\left[\left(l+yP_{x_1x_2}^i\right)^2yC^i\right]^4}},`$ (78) with the four-vector $`P_{x_1x_2}^i`$ defined by $$P_{x_1x_2}^i\left(qx_1q^{^{}}\right)x_2+k^{^{}},$$ (79) and the scalar $`C^i`$ defined by $$C^i\mathrm{\hspace{0.33em}2}k.q^{^{}}x_1x_2+y\left(P_{x_1x_2}^i\right)^2.$$ (80) The final result for the amplitude $`M_{V3}^i`$ is then given by $`M_{V3}^i=M_{BH}^i{\displaystyle \frac{e^2}{\left(4\pi \right)^2}}\{[{\displaystyle \frac{1}{\epsilon _{IR}}}\gamma _E+\mathrm{ln}\left({\displaystyle \frac{4\pi \mu ^2}{m^2}}\right)]{\displaystyle \frac{v^2+1}{v}}\mathrm{ln}\left({\displaystyle \frac{v+1}{v1}}\right)`$ (81) $`+{\displaystyle \frac{v^2+1}{2v}}\mathrm{ln}\left({\displaystyle \frac{v+1}{v1}}\right)\mathrm{ln}\left({\displaystyle \frac{v^21}{4v^2}}\right)+{\displaystyle \frac{v^2+1}{v}}[Sp\left({\displaystyle \frac{v+1}{2v}}\right)Sp\left({\displaystyle \frac{v1}{2v}}\right)]\}`$ (82) $`+{\displaystyle \frac{ie^5}{\left(4\pi \right)^2}}{\displaystyle \frac{1}{\left(p^{^{}}p\right)^2}}\overline{N}(p^{^{}},s_p^{^{}})\mathrm{\Gamma }_\nu (p^{^{}},p)N(p,s_p){\displaystyle \underset{0}{\overset{1}{}}}𝑑yy{\displaystyle \underset{0}{\overset{1}{}}}𝑑x_2x_2{\displaystyle \underset{0}{\overset{1}{}}}𝑑x_1`$ (83) $`\times \overline{u}(k^{^{}},h^{})\{\gamma ^\nu {\displaystyle \frac{\left(\mathit{}\mathit{}^{^{}}+m\right)}{2k.q^{^{}}}}\mathit{\epsilon ̸}^{}\mathrm{\hspace{0.17em}4}k.k^{^{}}[{\displaystyle \frac{2}{C^i}}+{\displaystyle \frac{1}{\left(C^i\right)^2}}(y\left(P_{x_1x_2}^i\right)^2+2P_{x_1x_2}^i.(kq^{^{}}))]`$ (84) $`{\displaystyle \frac{2}{C^i}}\left[\gamma ^\nu \mathit{\epsilon ̸}^{}\mathit{}^{^{}}+\mathit{}\gamma ^\nu \mathit{\epsilon ̸}^{}4m\epsilon ^\nu \gamma ^\nu \left(\mathit{}\mathit{}^{^{}}y\mathit{}_{x_1x_2}^i\right)\mathit{\epsilon ̸}^{}+y\mathit{\epsilon ̸}^{}\gamma ^\nu \mathit{}_{x_1x_2}^i+y\mathit{}_{x_1x_2}^i\mathit{\epsilon ̸}^{}\gamma ^\nu \right]`$ (85) $`+{\displaystyle \frac{1}{\left(C^i\right)^2}}[4(k.k^{^{}})\gamma ^\nu \mathit{}_{x_1x_2}^i\mathit{\epsilon ̸}^{}2\gamma ^\nu (\mathit{}\mathit{}^{^{}}y\mathit{}_{x_1x_2}^i+m)\mathit{\epsilon ̸}^{}\mathit{}_{x_1x_2}^i\mathit{}^{^{}}`$ (86) $`2\mathit{}\mathit{}_{x_1x_2}^i\gamma ^\nu \left(\mathit{}\mathit{}^{^{}}y\mathit{}_{x_1x_2}^i+m\right)\mathit{\epsilon ̸}^{}`$ (87) $`+y\mathit{}_{x_1x_2}^i(2\mathit{\epsilon ̸}^{}(\mathit{}\mathit{}^{^{}}y\mathit{}_{x_1x_2}^i)\gamma ^\nu +\mathrm{\hspace{0.17em}4}m\epsilon ^\nu )\mathit{}_{x_1x_2}^i]\}u(k,h).`$ (88) The Feynman parameter integrals in Eq. (81) will be performed numerically as explained in Section III C. In an analogous way, the result for the amplitude $`M_{V3}^f`$ corresponding to Fig. 2 (V3f) can be calculated, and yields as result : $`M_{V3}^f=M_{BH}^f{\displaystyle \frac{e^2}{\left(4\pi \right)^2}}\{[{\displaystyle \frac{1}{\epsilon _{IR}}}\gamma _E+\mathrm{ln}\left({\displaystyle \frac{4\pi \mu ^2}{m^2}}\right)]{\displaystyle \frac{v^2+1}{v}}\mathrm{ln}\left({\displaystyle \frac{v+1}{v1}}\right)`$ (89) $`+{\displaystyle \frac{v^2+1}{2v}}\mathrm{ln}\left({\displaystyle \frac{v+1}{v1}}\right)\mathrm{ln}\left({\displaystyle \frac{v^21}{4v^2}}\right)+{\displaystyle \frac{v^2+1}{v}}[Sp\left({\displaystyle \frac{v+1}{2v}}\right)Sp\left({\displaystyle \frac{v1}{2v}}\right)]\}`$ (90) $`+{\displaystyle \frac{ie^5}{\left(4\pi \right)^2}}{\displaystyle \frac{1}{\left(p^{^{}}p\right)^2}}\overline{N}(p^{^{}},s_p^{^{}})\mathrm{\Gamma }_\nu (p^{^{}},p)N(p,s_p){\displaystyle \underset{0}{\overset{1}{}}}𝑑yy{\displaystyle \underset{0}{\overset{1}{}}}𝑑x_2x_2{\displaystyle \underset{0}{\overset{1}{}}}𝑑x_1`$ (91) $`\times \overline{u}(k^{^{}},h^{})\{\mathit{\epsilon ̸}^{}{\displaystyle \frac{\left(\mathit{}^{^{}}+\mathit{}^{^{}}+m\right)}{2k^{^{}}.q^{^{}}}}\gamma ^\nu \mathrm{\hspace{0.17em}4}k.k^{^{}}[{\displaystyle \frac{2}{C^f}}+{\displaystyle \frac{1}{\left(C^f\right)^2}}(y\left(P_{x_1x_2}^f\right)^2+2P_{x_1x_2}^f.(k^{^{}}+q^{^{}}))]`$ (92) $`+{\displaystyle \frac{2}{C^f}}\left[\mathit{\epsilon ̸}^{}\gamma ^\nu \mathit{}^{^{}}+\mathit{}\mathit{\epsilon ̸}^{}\gamma ^\nu 4m\epsilon ^\nu \mathit{\epsilon ̸}^{}\left(\mathit{}^{^{}}+\mathit{}^{^{}}y\mathit{}_{x_1x_2}^f\right)\gamma ^\nu +y\gamma ^\nu \mathit{\epsilon ̸}^{}\mathit{}_{x_1x_2}^f+y\mathit{}_{x_1x_2}^f\gamma ^\nu \mathit{\epsilon ̸}^{}\right]`$ (93) $`+{\displaystyle \frac{1}{\left(C^f\right)^2}}[4(k.k^{^{}})\mathit{\epsilon ̸}^{}\mathit{}_{x_1x_2}^f\gamma ^\nu 2\mathit{\epsilon ̸}^{}(\mathit{}^{^{}}+\mathit{}^{^{}}y\mathit{}_{x_1x_2}^f+m)\gamma ^\nu \mathit{}_{x_1x_2}^f\mathit{}^{^{}}`$ (94) $`2\mathit{}\mathit{}_{x_1x_2}^f\mathit{\epsilon ̸}^{}\left(\mathit{}^{^{}}+\mathit{}^{^{}}y\mathit{}_{x_1x_2}^f+m\right)\gamma ^\nu `$ (95) $`+y\mathit{}_{x_1x_2}^f(2\gamma ^\nu (\mathit{}^{^{}}+\mathit{}^{^{}}y\mathit{}_{x_1x_2}^f)\mathit{\epsilon ̸}^{}+\mathrm{\hspace{0.17em}4}m\epsilon ^\nu )\mathit{}_{x_1x_2}^f]\}u(k,h),`$ (96) with the four-vector $`P_{x_1x_2}^f`$ defined by $$P_{x_1x_2}^f\left(qx_1q^{^{}}\right)x_2+k,$$ (97) and the scalar $`C^f`$ defined by $$C^f2k^{^{}}.q^{^{}}x_1x_2+y\left(P_{x_1x_2}^f\right)^2.$$ (98) Remark that in the vertex correction diagrams where the photon couples to the final electron (diagrams of Fig. 2 denoted by $`f`$), the invariant mass of the virtual $`(e^{}+\gamma ^{})`$ state in the loop is given by $`m^2+2k^{^{}}.q^{^{}}m^2`$. This means that an on-shell propagation is possible for the $`(e^{}+\gamma ^{})`$ state. This translates mathematically into the presence of integrable singularities in the corresponding Feynman parameter integrals of Eqs. (30, 52) and (89), and yields an imaginary part for the corresponding amplitude. In contrast, in the vertex correction diagrams where the photon couples to the initial electron (diagrams of Fig. 2 denoted by $`i`$), the invariant mass of the virtual $`(e^{}+\gamma ^{})`$ system in the loop is given by $`m^2`$ which means that the corresponding integrals contain no singularities. The numerical treatment of those singular Feynman parameter integrals will be discussed in section III C. #### 4 Electron self-energy diagrams of Figs. 2 (Si) and (Sf) We next evaluate the electron self-energy diagrams of Figs. 2 (Si) and (Sf). We only have to consider those diagrams where a photon is emitted and re-absorbed by an intermediate electron line. The diagrams with a loop on the initial or final electron lines are already absorbed in the wavefunction and electron mass renormalization, and therefore do not yield an additional correction. This can also be seen from the expression Eq. (A44) for the renormalized lepton self-energy, which vanishes on-shell. The amplitude corresponding to Fig. 2 (Si) is then given by : $`M_{Si}`$ $`=`$ $`ie^3{\displaystyle \frac{1}{\left(p^{^{}}p\right)^2}}\overline{N}(p^{^{}},s_p^{^{}})\mathrm{\Gamma }_\nu (p^{^{}},p)N(p,s_p)`$ (99) $`\times `$ $`\overline{u}(k^{^{}},h^{})\gamma ^\nu {\displaystyle \frac{\left(\mathit{}\mathit{}^{^{}}+m\right)}{2k.q^{^{}}}}\stackrel{~}{\mathrm{\Sigma }}(kq^{}){\displaystyle \frac{\left(\mathit{}\mathit{}^{^{}}+m\right)}{2k.q^{^{}}}}\mathit{\epsilon ̸}^{}u(k,h),`$ (100) where the renormalized self-energy is denoted by $`\stackrel{~}{\mathrm{\Sigma }}`$ and is given by Eq. (A44). Remark that the UV divergence in the loop integral of Fig. 2 (Si) has been removed through the renormalization of the electron field and electron mass. The UV finite renormalized self-energy $`\stackrel{~}{\mathrm{\Sigma }}`$ contains however an IR divergence from the counterterms. Inserting the expression for $`\stackrel{~}{\mathrm{\Sigma }}`$ (Eq. (A44)) into Eq.(99), yields : $`M_{Si}`$ $`=`$ $`M_{BH}^i{\displaystyle \frac{e^2}{(4\pi )^2}}\mathrm{\hspace{0.17em}2}\left[{\displaystyle \frac{1}{\epsilon _{IR}}}\gamma _E+\mathrm{ln}\left({\displaystyle \frac{4\pi \mu ^2}{m^2}}\right)\right]`$ (101) $`+`$ $`{\displaystyle \frac{ie^5}{(4\pi )^2}}{\displaystyle \frac{1}{\left(p^{^{}}p\right)^2}}\overline{N}(p^{^{}},s_p^{^{}})\mathrm{\Gamma }_\nu (p^{^{}},p)N(p,s_p)`$ (104) $`\times \overline{u}(k^{^{}},h^{})\gamma ^\nu {\displaystyle \frac{\left(\mathit{}\mathit{}^{^{}}+m\right)}{2k.q^{^{}}}}\{{\displaystyle \frac{m\left(\mathit{}\mathit{}^{}\right)}{m^22k.q^{}}}[1+{\displaystyle \frac{2m^2+6k.q^{}}{m^22k.q^{}}}\mathrm{ln}\left({\displaystyle \frac{2k.q^{}}{m^2}}\right)]`$ $`+[3{\displaystyle \frac{2m^2+2k.q^{}}{m^22k.q^{}}}\mathrm{ln}\left({\displaystyle \frac{2k.q^{}}{m^2}}\right)]\}\mathit{\epsilon ̸}^{}u(k,h).`$ The amplitude corresponding to Fig. 2 (Sf) is given by : $`M_{Sf}`$ $`=`$ $`ie^3{\displaystyle \frac{1}{\left(p^{^{}}p\right)^2}}\overline{N}(p^{^{}},s_p^{^{}})\mathrm{\Gamma }_\nu (p^{^{}},p)N(p,s_p)`$ (105) $`\times `$ $`\overline{u}(k^{^{}},h^{})\mathit{\epsilon ̸}^{}{\displaystyle \frac{\left(\mathit{}^{}+\mathit{}^{}+m\right)}{2k^{}.q^{}}}\stackrel{~}{\mathrm{\Sigma }}(k^{}+q^{}){\displaystyle \frac{\left(\mathit{}^{}+\mathit{}^{}+m\right)}{2k.q^{}}}\gamma ^\nu u(k,h),`$ (106) which can be worked out analogously as before and yields : $`M_{Sf}`$ $`=`$ $`M_{BH}^f{\displaystyle \frac{e^2}{(4\pi )^2}}\mathrm{\hspace{0.17em}2}\left[{\displaystyle \frac{1}{\epsilon _{IR}}}\gamma _E+\mathrm{ln}\left({\displaystyle \frac{4\pi \mu ^2}{m^2}}\right)\right]`$ (107) $`+`$ $`{\displaystyle \frac{ie^5}{(4\pi )^2}}{\displaystyle \frac{1}{\left(p^{^{}}p\right)^2}}\overline{N}(p^{^{}},s_p^{^{}})\mathrm{\Gamma }_\nu (p^{^{}},p)N(p,s_p)`$ (110) $`\times \overline{u}(k^{^{}},h^{})\mathit{\epsilon ̸}^{}{\displaystyle \frac{\left(\mathit{}^{}+\mathit{}^{}+m\right)}{2k^{}.q^{}}}\{{\displaystyle \frac{m\left(\mathit{}^{}+\mathit{}^{}\right)}{m^2+2k^{}.q^{}}}[1+{\displaystyle \frac{2m^26k^{}.q^{}}{m^2+2k^{}.q^{}}}\mathrm{ln}\left({\displaystyle \frac{2k^{}.q^{}}{m^2}}\right)]`$ $`+[3{\displaystyle \frac{2m^22k^{}.q^{}}{m^2+2k^{}.q^{}}}\mathrm{ln}\left({\displaystyle \frac{2k^{}.q^{}}{m^2}}\right)]\}\gamma ^\nu u(k,h).`$ Note that in Fig. 2 (Sf), the four-momentum squared of the $`(e^{}+\gamma ^{})`$ state in the loop is given by $`(k^{}+q^{})^2=m^2+2k^{}.q^{}m^2`$. Therefore, the self-energy and the amplitude for Fig. 2 (Sf) is complex, as was also noted for the vertex diagrams of Fig. 2 where the photon is emitted from the final electron (denoted by $`f`$). Eq. (107) yields indeed a complex amplitude because $`\mathrm{ln}(2k^{}.q^{}/m^2)=\mathrm{ln}(2k^{}.q^{}/m^2)+i\pi `$, for $`k^{}.q^{}>0`$. #### 5 Vertex correction diagram of Fig. 2 (V4) The vertex correction to the VCS process is given by Fig. 2 (V4), and its calculation is the same as the one for elastic electron scattering. This yields for the renormalized vertex correction : $$M_{V4}=ie^3\overline{u}(k^{^{}},h^{})\left[\left(F(Q^2)F(Q^2=0)\right)\gamma _\nu G(Q^2)i\sigma _{\nu \kappa }\frac{q^\kappa }{2m}\right]u(k,h)\frac{1}{q^2}\epsilon _\mu ^{}H^{\mu \nu }.$$ (111) In Eq. (111), $`F(Q^2)F(Q^2=0)`$ is given by Eq. (A23) and reduces in the ultrarelativistic limit ($`Q^2>>m^2`$) to Eq. (A26). The magnetic correction $`G(Q^2)`$ is given by Eq. (A14), and vanishes in the ultrarelativistic limit. #### 6 Vacuum polarization diagrams of Figs. 2 (P1i, P1f) and (P2) The vacuum polarization corrections of Figs. 2 (P1i, P1f) and (P2) involve the renormalized photon self-energy $`\stackrel{~}{\mathrm{\Pi }}(Q^2)`$, which has been calculated in appendix A. Therefore, we get for the vacuum polarization correction to the BH process (Figs. 2 (P1i, P1f)) : $$M_{P1}^i=M_{BH}^i\frac{1}{1\stackrel{~}{\mathrm{\Pi }}(t)},\mathrm{a}ndM_{P1}^f=M_{BH}^f\frac{1}{1\stackrel{~}{\mathrm{\Pi }}(t)},$$ (112) with $`t=(p^{}p)^2`$. Similarly, we get for the vacuum polarization correction to the VCS process (Fig. 2 (P2)) : $$M_{P2}=M_{VCS}\frac{1}{1\stackrel{~}{\mathrm{\Pi }}(Q^2)}.$$ (113) In the ultrarelativistic limit ($`Q^2>>m^2`$), $`\stackrel{~}{\mathrm{\Pi }}(Q^2)`$ is obtained from Eq. (A54) $$\stackrel{~}{\mathrm{\Pi }}(Q^2)=\frac{e^2}{(4\pi )^2}\frac{4}{3}\left\{\frac{5}{3}+\mathrm{ln}\left(\frac{Q^2}{m^2}\right)\right\}.$$ (114) ### B Soft-photon emission contributions and cancellation of IR divergences After removing the UV divergences from the virtual photon corrections to the $`epep\gamma `$ reaction in the last section, the resulting expressions still contain IR divergences. Both the corrections to the BH process of Figs. 2 (V1i, V1f, V2i, V2f, V3i, V3f, Si and Sf) and the vertex correction of Fig. 2 (V4) to the VCS process contain IR divergences. It is known for QED since a long time , that these IR divergences are cancelled at the cross section level by soft photon emission contributions. These soft photons are emitted from the charged particle lines and can have energies up to some maximal value $`\mathrm{\Delta }E_s`$ which is related to the finite resolution of the detector. In appendix A (section A 5), we calculate the soft bremsstrahlung contribution to electron scattering by performing the phase space integral over the soft photon in an exact way, and give the finite correction (after cancellation of all IR divergences) to the elastic electron scattering cross section. In this section, we generalize the result of appendix A to the case of the $`epep\gamma `$ reaction. The diagrams for the $`epep\gamma `$ reaction with one additional soft photon are shown in Fig. 3, where the hard photon of the $`epep\gamma `$ process is indicated by its four-momentum $`q^{}`$. In this section, we will show that the soft photon emission contributions of Fig. 3 contain IR divergences which exactly cancel the IR divergences appearing in the virtual photon correction diagrams of Fig. 2. The process where the energy $`\mathrm{\Delta }E_s`$ of the additionally emitted photon is not very small compared with the lepton momenta in the process, makes up the radiative tail to the $`epep\gamma `$ reaction. Its calculation will be discussed in section IV. #### 1 Factorization of amplitude for soft-photon emission processes Here, we evaluate the diagrams of Fig. 3 in the soft photon limit, i.e. when the second emitted photon has an energy much smaller than the initial and final lepton energies and also smaller than the hard photon (denoted by $`q^{}`$) in order to distinguish both photons. We will see that only the diagrams where a soft photon couples to an on-shell lepton contain IR divergences and lead to a finite logarithmic correction in $`\mathrm{\Delta }E_s`$. The amplitude corresponding with Fig. 3 (b1i) is given by : $`M_{b1i}=`$ $`ie^3\overline{u}(k^{^{}},h^{})\gamma ^\nu {\displaystyle \frac{\left(\mathit{}\mathit{}^{}\mathit{}+m\right)}{2k.q^{}2l.(kq^{})}}\mathit{\epsilon ̸}^{}(q^{}){\displaystyle \frac{\left(\mathit{}\mathit{}+m\right)}{2k.l}}\left(e\mathit{\epsilon ̸}^{}(l)\right)u(k,h)`$ (116) $`\times {\displaystyle \frac{1}{\left(p^{^{}}p\right)^2}}\overline{N}(p^{^{}},s_p^{^{}})\mathrm{\Gamma }_\nu (p^{^{}},p)N(p,s_p),`$ where $`l`$ is the four-momentum of the soft photon. In the soft photon limit ($`l0`$), Eq. (116) simplifies by using $`\left(\mathit{}\mathit{}+m\right)\gamma ^\alpha u(k,h)=\left(2k^\alpha \mathit{}\gamma ^\alpha \right)u(k,h)\mathrm{\hspace{0.17em}2}k^\alpha u(k,h)`$, which yields for Eq. (116) in the soft photon limit : $$M_{b1i}=M_{BH}^i(e)\epsilon _\alpha ^{}(l)\left[\frac{k^\alpha }{k.l}\right],$$ (117) where $`M_{BH}^i`$ is the Bethe-Heitler amplitude of Eq. (1) - corresponding with photon emission from the initial lepton. Similarly, we can derive the amplitude for Figs. 3 (b2i, b1f and b2f) which yields in the soft photon limit : $`M_{b1i}+M_{b2i}`$ $`=`$ $`M_{BH}^i(e)\epsilon _\alpha ^{}(l)\left[{\displaystyle \frac{k^{}_{}{}^{}\alpha }{k^{}.l}}{\displaystyle \frac{k^\alpha }{k.l}}\right],`$ (118) $`M_{b1f}+M_{b2f}`$ $`=`$ $`M_{BH}^f(e)\epsilon _\alpha ^{}(l)\left[{\displaystyle \frac{k^{}_{}{}^{}\alpha }{k^{}.l}}{\displaystyle \frac{k^\alpha }{k.l}}\right],`$ (119) where $`M_{BH}^f`$ is the Bethe-Heitler amplitude of Eq. (2) - corresponding with photon emission from the final lepton. Figs. 3 (b3i) and (b3f) contain the contributions where the soft photon couples to an off-shell lepton line. The amplitude corresponding with Fig. 3 (b3i) is given by : $`M_{b3i}=`$ $`ie^4\overline{u}(k^{^{}},h^{})\gamma ^\nu {\displaystyle \frac{\left(\mathit{}\mathit{}^{}\mathit{}+m\right)}{2k.q^{}2l.(kq^{})}}\mathit{\epsilon ̸}^{}(l){\displaystyle \frac{\left(\mathit{}\mathit{}^{}+m\right)}{2k.q^{}}}\mathit{\epsilon ̸}^{}(q^{})u(k,h)`$ (121) $`\times {\displaystyle \frac{1}{\left(p^{^{}}p\right)^2}}\overline{N}(p^{^{}},s_p^{^{}})\mathrm{\Gamma }_\nu (p^{^{}},p)N(p,s_p),`$ In the soft photon limit, Eq. (121) can be simplified by using : $`{\displaystyle \frac{\left(\mathit{}\mathit{}^{}\mathit{}+m\right)}{2k.q^{}2l.(kq^{})}}\gamma ^\alpha {\displaystyle \frac{\left(\mathit{}\mathit{}^{}+m\right)}{2k.q^{}}}`$ $``$ $`{\displaystyle \frac{\left(\mathit{}\mathit{}^{}+m\right)}{2k.q^{}}}\gamma ^\alpha {\displaystyle \frac{\left(\mathit{}\mathit{}^{}+m\right)}{2k.q^{}}}`$ (122) $`=`$ $`{\displaystyle \frac{\left(\mathit{}\mathit{}^{}+m\right)}{2k.q^{}}}{\displaystyle \frac{(kq^{})^\alpha }{k.q^{}}}{\displaystyle \frac{\gamma ^\alpha }{2k.q^{}}}.`$ (123) Consequently, the amplitude of Eq. (121) is given by : $`M_{b3i}`$ $`=`$ $`M_{BH}^i(e)\epsilon _\alpha ^{}(l){\displaystyle \frac{(kq^{})^\alpha }{k.q^{}}}`$ (124) $`+`$ $`ie^4\overline{u}(k^{^{}},h^{}){\displaystyle \frac{\gamma ^\nu \mathit{\epsilon ̸}^{}(l)\mathit{\epsilon ̸}^{}(q^{})}{2k.q^{}}}u(k,h){\displaystyle \frac{1}{\left(p^{^{}}p\right)^2}}\overline{N}(p^{^{}},s_p^{^{}})\mathrm{\Gamma }_\nu (p^{^{}},p)N(p,s_p).`$ (125) Similarly, the amplitude corresponding with Fig. 3 (b3f) is given by : $`M_{b3f}`$ $`=`$ $`M_{BH}^f(e)\epsilon _\alpha ^{}(l){\displaystyle \frac{(k^{}+q^{})^\alpha }{k^{}.q^{}}}`$ (126) $`+`$ $`ie^4\overline{u}(k^{^{}},h^{}){\displaystyle \frac{\mathit{\epsilon ̸}^{}(q^{})\mathit{\epsilon ̸}^{}(l)\gamma ^\nu }{2k^{}.q^{}}}u(k,h){\displaystyle \frac{1}{\left(p^{^{}}p\right)^2}}\overline{N}(p^{^{}},s_p^{^{}})\mathrm{\Gamma }_\nu (p^{^{}},p)N(p,s_p).`$ (127) In complete analogy to Eqs. (118,119), we can also calculate the soft photon emission contributions to the VCS process. They are shown in Figs. 3 (b4) and (b5), and their calculation in the soft photon limit yields : $$M_{b4}+M_{b5}=M_{VCS}(e)\epsilon _\alpha ^{}(l)\left[\frac{k^{}_{}{}^{}\alpha }{k^{}.l}\frac{k^\alpha }{k.l}\right],$$ (128) where $`M_{VCS}`$ is the VCS amplitude of Eq. (4). We see from Eqs. (118,119 and 128), that for the diagrams of Fig. 3 where the soft photon couples to an on-shell lepton, the original amplitude factorizes : in Eqs. (118,119) the BH amplitude factorizes, and in Eq. (128) the VCS amplitude factorizes. The resulting amplitudes are proportional to $`1/l`$, which leads to a logarithmic divergence when integrating over the phase space of the soft photon. In contrast, the amplitudes of Eqs. (124,126) where the photon couples to an off-shell lepton line are finite when $`l0`$, and the corresponding phase space integral becomes vanishingly small in the limit $`l0`$. #### 2 Radiative correction due to soft-photon emission processes In the soft-photon limit, we therefore need only to keep the bremsstrahlung corrections of Eqs. (118,119 and 128), where the BH and VCS amplitudes factorize. To first order in $`\alpha _{em}`$ (relative to the BH + VCS cross section) the bremsstrahlung correction therefore amounts to calculate the phase space integral of the form : $`d\sigma {\displaystyle \frac{d^3\stackrel{}{k}_e^{^{}}}{\left(2\pi \right)^3\mathrm{\hspace{0.17em}2}E_e^{}}}{\displaystyle \frac{d^3\stackrel{}{q}^{^{}}}{\left(2\pi \right)^3\mathrm{\hspace{0.17em}2}|\stackrel{}{q}^{^{}}|}}`$ $`{\displaystyle \frac{d^3\stackrel{}{p}_N^{^{}}}{\left(2\pi \right)^3\mathrm{\hspace{0.17em}2}E_N^{}}}{\displaystyle \frac{d^3\stackrel{}{l}}{\left(2\pi \right)^3\mathrm{\hspace{0.17em}2}\mathrm{l}}}(2\pi )^4\delta ^4(k+pk^{}q^{}p^{}l)`$ (130) $`\times |M_{BH}+M_{VCS}|^2(e^2)[{\displaystyle \frac{k_\mu ^{^{}}}{k^{^{}}.l}}{\displaystyle \frac{k_\mu }{k.l}}].[{\displaystyle \frac{k^\mu }{k^{^{}}.l}}{\displaystyle \frac{k^\mu }{k.l}}],`$ where $`\mathrm{l}|\stackrel{}{l}|`$ denotes the soft photon energy, and where the total BH amplitude is given by $`M_{BH}=M_{BH}^i+M_{BH}^f`$. The calculation of the bremsstrahlung integral of Eq. (130) goes along similar lines as the corresponding integral for elastic scattering, for which the technical details can be found in appendix A (section A 5). We will point out in this section the differences which arise for the $`epep\gamma `$ reaction. There are two practical ways to measure the $`epep\gamma `$ reaction, by measuring two particles in the final state. One can either measure the outgoing electron in coincidence with the recoiling nucleon : this is the ideal technique when measuring the $`epep\gamma `$ reaction at low outgoing photon energy as is done in . The alternative is to measure the outgoing electron in coincidence with the photon : this is the technique when doing a very inelastic experiment, such as deeply virtual Compton scattering, where the photon is produced with a large energy. We discuss here first the case where one detects the outgoing electron and photon, and indicate at the end the changes which apply when measuring the outgoing electron and recoiling nucleon. If one measures the $`epep\gamma `$ reaction by detecting the outgoing electron and photon, one eliminates in Eq. (A56) the integral over $`\stackrel{}{p}_N^{^{}}`$ with the momentum conserving $`\delta `$-function, which gives : $`d\sigma `$ $`{\displaystyle \frac{d^3\stackrel{}{k}_e^{^{}}}{\left(2\pi \right)^3\mathrm{\hspace{0.17em}2}E_e^{}}}{\displaystyle \frac{d^3\stackrel{}{q}^{^{}}}{\left(2\pi \right)^3\mathrm{\hspace{0.17em}2}|\stackrel{}{q}^{^{}}|}}{\displaystyle \frac{d^3\stackrel{}{l}}{\left(2\pi \right)^3\mathrm{\hspace{0.17em}2}\mathrm{l}}}{\displaystyle \frac{1}{2E_N^{}}}`$ (133) $`\times (2\pi )\delta \left(E_e+E_NE_e^{}|\stackrel{}{q}^{^{}}|\sqrt{(\stackrel{}{q}+\stackrel{}{p}_N\stackrel{}{q}^{^{}}\stackrel{}{l})^2+M_N^2}\mathrm{l}\right)`$ $`\times |M_{BH}+M_{VCS}|^2(e^2)[{\displaystyle \frac{k_\mu ^{^{}}}{k^{^{}}.l}}{\displaystyle \frac{k_\mu }{k.l}}].[{\displaystyle \frac{k^\mu }{k^{^{}}.l}}{\displaystyle \frac{k^\mu }{k.l}}].`$ Due to the energy conserving $`\delta `$-function in Eq. (A58), the upper limit in the integration over the soft photon phase space depends on the angle. Therefore, this integration volume has a complicated ellipsoidal shape in the lab system. In order for the soft-photon phase space integration volume to be spherical, one has to perform the calculation in the c.m. system $`𝒮_1`$ of the (recoiling nucleon + soft-photon), generalizing the procedure of appendix A for elastic scattering to the $`epep\gamma `$ reaction. The system $`𝒮_1`$ is defined by : $`\stackrel{}{p}_N^{^{}}+\stackrel{}{l}=\stackrel{}{p}_N+\stackrel{}{q}\stackrel{}{q}^{^{}}=0`$. In the system $`𝒮_1`$, the energy conserving delta function in Eq. (133) is independent of the soft-photon angles, and the maximal soft photon energy is isotropic. The integral over the soft-photon momentum (up to some maximum value $`\mathrm{\Delta }E_s`$) can then be performed independently from the integration over the soft photon emission angles. If $`\mathrm{\Delta }E_s`$ is sufficiently small, one can furthermore neglect the soft photon energy with respect to the other energies in the $`\delta `$-function, and perform the integral over the photon momentum $`|\stackrel{}{q}^{^{}}|`$ in Eq. (133) to obtain the correction to the fivefold differential $`epep\gamma `$ cross section. We indicate in the following only how the squared matrix element for the $`epep\gamma `$ reaction is modified due to soft photon emission. This correction due to soft bremsstrahlung is given by : $$|M_{epep\gamma }^{SOFT\gamma }|^2=|M_{BH}+M_{VCS}|^2\left(e^2\right)\frac{d^3\stackrel{}{l}}{\left(2\pi \right)^3\mathrm{\hspace{0.17em}2}\mathrm{l}}\left[\frac{k_\mu ^{^{}}}{k^{^{}}.l}\frac{k_\mu }{k.l}\right].\left[\frac{k^\mu }{k^{^{}}.l}\frac{k^\mu }{k.l}\right].$$ (134) The correction factor multiplying $`|M_{BH}+M_{VCS}|^2`$ gives immediately the correction factor to the fivefold $`epep\gamma `$ cross section. In Eq. (134), the soft-photon phase space integral is understood to be performed in the system $`𝒮_1`$, where the integration volume is spherical. Its calculation was already performed in appendix A. One sees that the integral in Eq. (134) has a logarithmic IR divergence, corresponding with the emission of photons with zero energy. To evaluate it, one has to regularize it, which is done in this work by using dimensional regularization. This amounts to evaluate the integral (in the system $`𝒮_1`$) in $`D1`$ dimensions ($`D4`$ corresponds to the physical limit). This calculation is performed in appendix A and yields (similar to Eq. (A95)) as result : $`|M_{epep\gamma }^{SOFT\gamma }|^2`$ $`=`$ $`|M_{BH}+M_{VCS}|^2`$ (135) $`\times `$ $`\left\{{\displaystyle \frac{e^2}{4\pi ^2}}\left[{\displaystyle \frac{1}{\epsilon _{IR}}}+\gamma _E\mathrm{ln}\left({\displaystyle \frac{4\pi \mu ^2}{m^2}}\right)\right]\left[{\displaystyle \frac{v^2+1}{2v}}\mathrm{ln}\left({\displaystyle \frac{v+1}{v1}}\right)\mathrm{\hspace{0.17em}1}\right]+\delta _R\right\},`$ (136) In Eq. (136), $`\delta _R`$ is the finite part of the real radiative correction corresponding with soft photon emission, and is given as in appendix A (Eq. (A102)) by : $`\delta _R`$ $`\stackrel{Q^2>>m^2}{}`$ $`{\displaystyle \frac{\alpha _{em}}{\pi }}\{\mathrm{ln}\left({\displaystyle \frac{(\mathrm{\Delta }E_s)^2}{\stackrel{~}{E}_e\stackrel{~}{E}_e^{^{}}}}\right)[\mathrm{ln}\left({\displaystyle \frac{Q^2}{m^2}}\right)\mathrm{\hspace{0.17em}1}]`$ (138) $`{\displaystyle \frac{1}{2}}\mathrm{ln}^2\left({\displaystyle \frac{\stackrel{~}{E}_e}{\stackrel{~}{E}_e^{^{}}}}\right)+{\displaystyle \frac{1}{2}}\mathrm{ln}^2\left({\displaystyle \frac{Q^2}{m^2}}\right){\displaystyle \frac{\pi ^2}{3}}+Sp\left(\mathrm{cos}^2{\displaystyle \frac{\stackrel{~}{\theta }_e}{2}}\right)\},`$ In Eq. (138), we next have to express the kinematical variables ($`\stackrel{~}{E}_e,\stackrel{~}{E}_e^{},\mathrm{cos}\stackrel{~}{\theta }_e`$) in the system $`𝒮_1`$ (denoted by tilded quantities) in terms of the lab quantities, which we denote by untilded quantities ($`E_e,E_e^{},\mathrm{cos}\theta _e`$). To make the transformation between the system $`𝒮_1`$ and the lab system, we first introduce the missing four-momentum $`p_{m1}p_N^{^{}}+l`$. The system $`𝒮_1`$ is defined by $`\stackrel{}{p}_{m1}=\stackrel{}{0}`$, and the missing mass $`M_{m1}`$ of the system ($`p^{}+l`$) is defined by : $$M_{m1}^2=(p^{}+l)^2=(p+qq^{})^2.$$ (139) We can then easily express the electron energies and angle in the system $`𝒮_1`$ in terms of lab quantities : $`\stackrel{~}{E}_e={\displaystyle \frac{k.p_{m1}}{M_{m1}}}={\displaystyle \frac{1}{M_{m1}}}k.(p+qq^{})={\displaystyle \frac{M_N}{M_{m1}}}\left(E_e{\displaystyle \frac{Q^2}{2M_N}}{\displaystyle \frac{kq^{}}{M_N}}\right),`$ (140) $`\stackrel{~}{E}_e^{}={\displaystyle \frac{k^{}.p_{m1}}{M_{m1}}}={\displaystyle \frac{1}{M_{m1}}}k^{}.(p+qq^{})={\displaystyle \frac{M_N}{M_{m1}}}\left(E_e^{}+{\displaystyle \frac{Q^2}{2M_N}}{\displaystyle \frac{k^{}q^{}}{M_N}}\right),`$ (141) $`\mathrm{sin}^2\stackrel{~}{\theta }_e/2={\displaystyle \frac{E_eE_e^{}}{\stackrel{~}{E}_e\stackrel{~}{E}_e^{}}}\mathrm{sin}^2\theta _e/2.`$ (142) The maximal soft-photon energy $`\mathrm{\Delta }E_s`$ in the system $`𝒮_1`$, is given by : $$\mathrm{\Delta }E_s=\frac{M_{m1}^2M_N^2}{2M_{m1}}.$$ (143) If one measures the $`epep\gamma `$ reaction by detecting the outgoing electron and recoiling proton, the derivation goes along similar lines as above. One starts now by eliminating in Eq. (A56) the integral over $`\stackrel{}{q}^{^{}}`$. Then one goes into the c.m. system $`𝒮_2`$ of the (VCS photon $`q^{}`$ \+ soft photon), where the energy conserving $`\delta `$-function is independent of the soft-photon angles, and where the maximal soft photon energy is isotropic. This system $`𝒮_2`$ is defined by : $`\stackrel{}{q}^{^{}}+\stackrel{}{l}=\stackrel{}{p}_N+\stackrel{}{q}\stackrel{}{p}_N^{^{}}=0`$. The calculation of the soft-photon emission integral is then completely similar as above, and leads to the finite correction of Eq.(138), where the the kinematical variables ($`\stackrel{~}{E}_e,\stackrel{~}{E}_e^{},\mathrm{cos}\stackrel{~}{\theta }_e`$) are now understood in the system $`𝒮_2`$. To make the transformation between the system $`𝒮_2`$ and the lab system, we first introduce the missing four-momentum $`p_{m2}q^{^{}}+l`$. The system $`𝒮_2`$ is defined by $`\stackrel{}{p}_{m2}=\stackrel{}{0}`$, and the missing mass $`M_{m2}`$ of the system ($`q^{}+l`$) is defined by : $$M_{m2}^2=(q^{}+l)^2=(p+qp^{})^2.$$ (144) We can then easily express the electron energies and angle in the system $`𝒮_2`$ in terms of lab quantities : $`\stackrel{~}{E}_e={\displaystyle \frac{k.p_{m2}}{M_{m2}}}={\displaystyle \frac{1}{M_{m2}}}k.(p+qp^{})={\displaystyle \frac{M_N}{M_{m2}}}\left(E_e{\displaystyle \frac{Q^2}{2M_N}}{\displaystyle \frac{kp^{}}{M_N}}\right),`$ (145) $`\stackrel{~}{E}_e^{}={\displaystyle \frac{k^{}.p_{m2}}{M_{m2}}}={\displaystyle \frac{1}{M_{m2}}}k^{}.(p+qp^{})={\displaystyle \frac{M_N}{M_{m2}}}\left(E_e^{}+{\displaystyle \frac{Q^2}{2M_N}}{\displaystyle \frac{k^{}p^{}}{M_N}}\right),`$ (146) $`\mathrm{sin}^2\stackrel{~}{\theta }_e/2={\displaystyle \frac{E_eE_e^{}}{\stackrel{~}{E}_e\stackrel{~}{E}_e^{}}}\mathrm{sin}^2\theta _e/2.`$ (147) The maximal soft-photon energy $`\mathrm{\Delta }E_s`$ in the system $`𝒮_2`$, is given by : $$\mathrm{\Delta }E_s=\frac{M_{m2}}{2},$$ (148) #### 3 Cancellation of IR divergences We can now demonstrate for the $`epep\gamma `$ reaction, that the IR divergences from the soft photon emission corrections exactly cancel against the IR divergences from the virtual radiative corrections, calculated in section III A. Concentrating here only on the IR divergent parts of the virtual radiative corrections, we found in section III A that the amplitudes of Eqs. (19,30,43,52,81,89,101,107, and 111) contain IR divergences. Those IR divergent parts are given by : $`M_{V1}^i+(CT)_{V1}^i+M_{V1}^f+(CT)_{V1}^fM_{BH}{\displaystyle \frac{e^2}{4\pi ^2}}\left({\displaystyle \frac{1}{2}}\right)\left[{\displaystyle \frac{1}{\epsilon _{IR}}}\gamma _E+\mathrm{ln}\left({\displaystyle \frac{4\pi \mu ^2}{m^2}}\right)\right],`$ (149) $`M_{V2}^i+(CT)_{V2}^i+M_{V2}^f+(CT)_{V2}^fM_{BH}{\displaystyle \frac{e^2}{4\pi ^2}}\left({\displaystyle \frac{1}{2}}\right)\left[{\displaystyle \frac{1}{\epsilon _{IR}}}\gamma _E+\mathrm{ln}\left({\displaystyle \frac{4\pi \mu ^2}{m^2}}\right)\right],`$ (150) $`M_{V3}^i+M_{V3}^fM_{BH}{\displaystyle \frac{e^2}{4\pi ^2}}\left({\displaystyle \frac{1}{2}}\right){\displaystyle \frac{v^2+1}{2v}}\mathrm{ln}\left({\displaystyle \frac{v+1}{v1}}\right)\left[{\displaystyle \frac{1}{\epsilon _{IR}}}\gamma _E+\mathrm{ln}\left({\displaystyle \frac{4\pi \mu ^2}{m^2}}\right)\right],`$ (151) $`M_{Si}+M_{Sf}M_{BH}{\displaystyle \frac{e^2}{4\pi ^2}}\left({\displaystyle \frac{1}{2}}\right)\left[{\displaystyle \frac{1}{\epsilon _{IR}}}\gamma _E+\mathrm{ln}\left({\displaystyle \frac{4\pi \mu ^2}{m^2}}\right)\right],`$ (152) $`M_{V4}+(CT)_{V4}M_{VCS}{\displaystyle \frac{e^2}{4\pi ^2}}\left({\displaystyle \frac{1}{2}}\right)\left[{\displaystyle \frac{v^2+1}{2v}}\mathrm{ln}\left({\displaystyle \frac{v+1}{v1}}\right)1\right]\left[{\displaystyle \frac{1}{\epsilon _{IR}}}\gamma _E+\mathrm{ln}\left({\displaystyle \frac{4\pi \mu ^2}{m^2}}\right)\right].`$ (153) Adding them all up gives the following correction to the squared amplitude for the virtual radiative corrections : $`|M_{BH}+M_{DVCS}+M_{epep\gamma }^{VIRTUAL\gamma }|^2`$ (154) $`=|M_{BH}+M_{VCS}|^2\left\{1+{\displaystyle \frac{e^2}{4\pi ^2}}\left[{\displaystyle \frac{v^2+1}{2v}}\mathrm{ln}\left({\displaystyle \frac{v+1}{v1}}\right)1\right]\left[{\displaystyle \frac{1}{\epsilon _{IR}}}\gamma _E+\mathrm{ln}\left({\displaystyle \frac{4\pi \mu ^2}{m^2}}\right)\right]\right\}+\mathrm{},`$ (155) where the ellipses denote the finite first order virtual radiative correction to the $`epep\gamma `$ reaction as was calculated and can be found in section III A. Adding the virtual (Eq.(155)) and real (Eq.(136)) radiative corrections to the $`epep\gamma `$ reaction, one verifies that the IR divergences in the sum exactly cancel, showing QED at work! Note that this cancellation is different as compared to the case of elastic electron scattering. Indeed, for the virtual photon correction diagrams to the Bethe Heitler process, there are 3 types of vertex diagrams (Eqs. (149,150 and 151)) and the self energy diagram (Eq. (152)), and the corresponding counterterms, which have an IR divergence. On the other hand, for the virtual radiative corrections to elastic electron scattering, there is only one vertex diagram which is IR divergent. ### C Integration method for the virtual photon corrections At this stage of the calculation of the first order QED radiative corrections to the $`epep\gamma `$ reaction, the treatment of all UV and IR divergences, resulting from the radiative corrections at the electron side, has been performed. The UV divergences have been removed by the renormalization procedure whereas the IR divergences were shown to cancel at the cross section level when adding the soft photon emission processes. Now, the evaluation of the remaining Feynman parameter integrals in the finite terms such as in Eq. (19) still has to be done. Among the one-loop virtual radiative corrections to the $`epep\gamma `$ reaction shown in Fig. 2, six give rise to simple analytical formulas. For the six vertex diagrams, denoted by V1i,V2i,V3i,V1f,V2f and V3f, the trick consisting of adding and subtracting the divergent term for each of them (as explained in Section III A) gives rise to Feynman parameter integrals that are rather complicated to be done analytically. Therefore, we will evaluate them in this work by a numerical procedure. Although these Feynman parameter integrals are by construction finite, appropriate numerical methods are needed to perform them. Two main difficulties are encountered in these numerical integrations. Firstly, the variations of the integrated functions are always extremely sharp near the integration limits. In fact, a typical behaviour is a rather flat dependence in the middle of the domain and two pronounced rises when approaching $`0`$ or $`1`$ for the Feynman parameters with a width of the order $`m/E_e`$. The contribution of these two peaks has to be evaluated carefully in order to obtain a good precision for the final result. Secondly, we know that the virtual radiative corrections to the $`epep\gamma `$ reaction allow the propagation of on-shell states (see section III A). This is mathematically expressed by the presence of integrable singularities in the Feynman parameter integrals which require an analytical continuation into the complex plane and gives rise to an imaginary part for the amplitude. To evaluate the Feynman parameter integrals, our strategy is to perform the first integration analytically. The last integrations will then be performed numerically using the Gauss-Legendre integration routine. The analytical calculation of the first integration provides a shorter calculational time and a higher precision. The main advantage however is that in the case of a singularity, the pole is avoided by deforming the integration contour into the complex plane, using analytical continuation. In this way, one removes the difficulties for the remaining integrations along the real axis. To classify the Feynman parameter integrals that occur in the six vertex diagrams under study, we start by factorizing all the Dirac $`\gamma `$ matrices and decomposing the components of the four-vectors. All resulting integrals then reduce to the generic form : $$\underset{0}{\overset{1}{}}𝑑x_1𝑑x_2𝑑x_3\frac{P(x_1,x_2,x_3)}{Q(x_1,x_2,x_3)},$$ (156) where $`P`$ and $`Q`$ are polynomials in three (real) Feynman parameters $`x_1,x_2,x_3`$. Let’s choose $`x_1`$ to be the more internal variable. Then the first integration is either of the form : $$\underset{0}{\overset{1}{}}\frac{x_1^mdx_1}{(\alpha x_1+\beta )^n},$$ (157) or $$\underset{0}{\overset{1}{}}\frac{x_1^mdx_1}{(\alpha x_1^2+\beta x_1+\gamma )^n},$$ (158) where $`\alpha `$, $`\beta `$ and $`\gamma `$ are polynomials in $`x_2`$ and $`x_3`$ with coefficients that are functions of kinematical variables. In Eqs. (157,158), $`m`$ varies from $`0`$ to $`4`$ and $`n`$ is equal to $`1`$ or $`2`$, to accomodate all cases appearing in section III A. These successive decompositions increase the number of terms to calculate but they have the advantage to provide two simple classes of integrals without any vector or matrix dependence. The possibility of poles in the integrands of Eqs. (157,158) naturally splits the problem into two parts, whether the integrand is regular or singular. #### 1 Regular integrand When the denominator doesn’t have any singularities, some recurrence relations exist for these integrals and can be found in Ref.. Unfortunately for small values of $`\alpha `$ as compared to $`\beta `$ or to $`\gamma `$, it has been seen that these relations are numerically unstable. This has thus led us to use several methods of integration with each a different domain of validity. For small ratios $`r`$ ($`r=\alpha /\beta `$ for Eq. (157) or $`r=\alpha /\gamma `$ for Eq. (158) ) as compared to 1, we perform a Taylor expansion of the integral and tune the order of each development to complete a fixed criterium of convergence (for example we require that the ratio between the last and the first terms is of the order of the numerical precision in double precision). For $`r>1`$ the recurrence relations are used as they are stable in this range. In the intermediate zone $`(0.2r1)`$, we use the Gauss-Legendre numerical integration method. #### 2 Singular integrand In the case of the propagation of on-shell intermediate states, the polynomials of the denominators in Eqs. (157,158) acquire one (or two) roots in the domain of integration. Some simple physical considerations have shown that among the six diagrams numerically evaluated, the three processes where the photon in the $`epep\gamma `$ reaction is emitted from the initial electron line are free of poles (section III A). In contrast, the three vertex graphs where the photon is emitted from the final electron line were seen to contain singularities. The corresponding integrals are then defined by an analytical continuation into the complex plane and take the form : $$\underset{0}{\overset{1}{}}\frac{x_1^mdx_1}{(\alpha x_1+\beta \pm iϵ)^n}\mathrm{o}r\underset{0}{\overset{1}{}}\frac{x_1^mdx_1}{(\alpha x_1^2+\beta x_1+\gamma \pm iϵ)^n}.$$ (159) The prescription for on-shell propagation is of course already taken into account in the propagators and determines the sign in front of the $`iϵ`$ (which can also be obtained by applying the simple trick $`mmiϵ/2`$). Complications can occur from the possibility of two distinct roots in the interval $`[0,1]`$ for the second order polynomial. An important remark then concerns the variable of integration. In Eq. (156), the choice of $`x_1`$ as the more internal dimension was purely arbitrary. In fact all the decompositions in the three parameters have been derived and it has been shown that it was always possible to find an expansion providing at most one singularity. In appendix B, we give the analytical results for the integrals of Eq. (159). We checked these results with a numerical method, where one pole along the interval is avoided by analytically continuing the integrand into the complex plane. In this way, the integral along is replaced by an integration along a semi-circle (with origin at $`0.5+0i`$ and radius 1/2) in the opposite complex half-plane with respect to the pole. A comparison between the two methods shows a perfect agreement. Only in the special cases where one pole comes close to an edge of the domain of integration (typically within a distance $`m^2/E_e^2`$ to 0 or 1), one has to increase the number of integration points of the numerical method to obtain the same precision. #### 3 Numerical checks and accuracy Thanks to the analytical calculation of the first integration in the Feynman parameter integrals under study, singularities on the real axis have been removed and the two remaining integrations can then be performed numerically using the Gauss-Legendre method. In the implementation of this algorithm the major difficulty consisted in finding the suitable binnig of the integration domain and to determine the number of points per bins. A detailed study of the integrated functions has been performed to estimate the width and amplitude of the sharp variations close to the ends of the domain. In this paragraph we discuss various checks of the precision of our results as well as their numerical stability. A strong cross check of the reliability of our calculations is the exact agreement between two programs developed in paralell . Both of them use the same numerical method but they have been coded independently using in most cases different decomposition of the terms and different order in the integration variables, which checks the symmetry in the permutation of $`x_1,x_2`$ and $`x_3`$ variables. Comparison at each intermediate stage of the calculation also excludes any missprints in the writing of the quite extensive expressions. Besides this agreement between two independent programs, the next requirement is the numerical convergence of the calculations. Figs. 4 and 5 show results obtained for typical MAMI and JLab kinematics respectively. Beyond a certain density of integration bins and points per bin, the numerical instabilities are brought down to few $`10^4`$ of the lowest order cross section. This accuracy is far below all the other theoretical uncertainties related to the performed approximations or experimental knowledge of the form factors (of the order of 1%). Nevertheless this kind of very good convergence is useful since numerical instabilities can be amplified in the coherent sum of all the diagrams or when computing higher energy kinematics. In the case of the deeply virtual Compton scattering, we have checked that one has to double the number of integration points, to get the same numerical precision. Some features of the electromagnetic interaction itself can also be used to check further the validity of our results. Let’s consider the total amplitude of the sum of all the virtual radiative correction diagrams (Fig. 2). Denoting the Lorentz index associated with the real photon vertex by $`\mu `$, this amplitude can be written as the scalar product $`T^\mu ϵ_\mu ^{}`$ where $`ϵ^{}`$ stands for the polarization vector of the real photon with four-vector $`q^{^{}}`$ and where $`T^\mu `$ represents the electromagnetic current. The gauge invariance of electromagnetism implies $`T^\mu q_\mu ^{^{}}=0`$ and provides us with a powerful test of our calculations. Since our numerical accuracy is finite, we cannot get exactly zero. Therefore, we rather define a quantity compared to which the scalar product $`T^\mu q_\mu ^{^{}}`$ has to be small. A natural quantity is the product of the norms of the two Lorentz vectors. The gauge invariance criteria thus becomes a test of the smallness of the following dimensionless ratio : $$\frac{\left|T^\mu q_\mu ^{^{}}\right|^2}{\left|T^\mu T_\mu ^{}\right|(q^{}_{}{}^{}0)^2}1$$ (160) This ratio is shown in Fig. 6 as a function of the angle between $`q`$ and $`q^{^{}}`$. The gauge invariance is verified by the fact that the smallest ratio (solid curve) stays in the range $`[10^4,10^6]`$ and is obtained when the complete set of diagrams with analytical+numerical terms is included in $`T^\mu `$. As a last consistency check, we investigated the mass dependence of the virtual radiative corrections. The relative effect in the BH + Born cross section is illustrated in Fig. 7 for different values of the mass of the lepton. For this test we kept track of the mass dependence in all the kinematical variables. We observe that when increasing the lepton mass (at fixed lepton kinematics), the effect of the radiative corrections rapidly decreases, which reflects the suppression of photon emission by a heavy particle. ### D Radiative corrections at the proton side and two-photon exchange corrections In section III A \- III C, we calculated the radiative corrections to the $`epep\gamma `$ reaction, corresponding with the diagrams of Figs. 2 and 3. They are the virtual radiative corrections at the lepton side, the vacuum polarization contributions and the soft-photon emission from the lepton. These can be calculated model-independently as has been shown above. Although these corrections are the dominant ones (when $`Q^2>>m^2`$, leading to large logarithms), we want to estimate in this section how large are the virtual radiative corrections at the proton side, the two-photon exchange corrections (direct and crossed box diagrams) and the soft-photon emission from the proton. Generally, the radiative corrections from the proton side are typically suppressed compared with those from the electron, due to the much larger mass of the proton. However, to calculate the first order radiative corrections to the $`epep\gamma `$ reaction which originate from the proton side, one needs a model for the VCS process. We do not aim in this paper, to calculate these corrections within a given model. However, to provide some quantitative estimate, we will follow the results of , where the corrections at the proton side were studied for elastic scattering. The $`Z`$-dependent corrections originate from the interference between soft-photon emission from the electron and from the proton, and from the two-photon exchange contributions (direct and crossed box diagrams). Both processes contain IR divergences, which cancel in their sum at the cross section level. The interference between the soft-photon emission from the electron and from the proton can be calculated along the same lines as in appendix A 5 for the electron (neglecting form factor effects in the soft-photon limit). For the two-photon exchange contributions, the calculation is dominated by those regions in the integration where one of the two exchanged photons is soft. Therefore, one can evaluate the rest of this amplitude by taking the momentum of either of the two exchanged photons to be zero. In this approximation, the original amplitude factorizes and one can follow the derivation of , where this same calculation has been performed for elastic scattering. Therefore, the $`Z`$-dependent radiative corrections can be estimated in the soft-photon limit by the same correction factor of Eq. (A116) as for elastic scattering. The $`Z^2`$-dependent corrections originate from the soft-bremsstrahlung from the proton and from the proton vertex corrections. In , these corrections have also been calculated for elastic scattering. For the soft-photon emission, one can again factorize the orginal amplitude, so that the same correction factor is obtained for the $`epep\gamma `$ reaction as for elastic scattering. The proton vertex correction has been split in into two parts. The first part contains entirely the IR divergence, which cancels with the IR divergence from soft-photon emission from the proton, and in which the original amplitude factorizes. The second term in the proton vertex correction depends on the nucleon structure (form factor dependence for elastic scattering) and will be different when going from elastic scattering to the $`epep\gamma `$ reaction. For elastic scattering, this structure dependent term was however found to be quite small, except when going to very large $`Q^2`$ (much larger than $`M_N^2`$). When staying in the few GeV<sup>2</sup> region, this correction was calculated in to be well below 1%. Therefore, we approximate the $`Z^2`$ dependent correction to the $`epep\gamma `$ reaction by the structure-independent term of Eq. (A119), as calculated in , and will neglect in the following the structure dependent term. ## IV Radiative tail for elastic scattering and VCS Besides the knowledge of the virtual radiative corrections and the soft-photon emission contributions to the $`epep\gamma `$ reaction, which were studied in section III, the accurate determination of the $`epep\gamma `$ cross section from measured spectra also implies the knowledge of the radiative tail. The radiative tail consists of the photon emission processes where a semi-hard photon (with energy not very small compared with e.g. the lepton energies) is radiated from the electron (or proton). The radiative tail to elastic or inelastic lepton-nucleon scattering has been the subject of numerous studies in the literature . The elastic radiative tail also makes a sizeable contribution to the cross sections for deep-inelastic lepton-nucleon scattering (see e.g. ). One should notice that the distinction between the soft-photon emission and the radiative tail is not a fundamental one, the latter being just the extension of photon emission processes to higher energies. Although the formulas given in this paper for the real radiative corrections can in principle be extended and applied to higher energies (e.g. Eqs. (A111-A114) for elastic scattering), in some cases the characteristics of the experimental detection apparatus can be such that the cut in $`E_e^{el}E_e^{}`$ (elastic case) or in the missing mass $`M_x^2`$ ($`M_{m1}^2`$ or $`M_{m2}^2`$ for the VCS case) cannot be cleanly defined, because the apparatus can have a changing acceptance as a function of $`E_e^{el}E_e^{}`$ or $`M_x^2`$, introducing a bias in the radiative tail. Therefore, it is useful to consider the radiative tail separately and to generate it in a Monte Carlo simulation. In doing such a simulation, it can be very helpful to have a “recipe”, because it is a way to fold radiative effects with acceptance functions and other effects (e.g. multiple scattering, energy loss by collision, external radiative effects). In the literature such “recipes” were quite often presented. Many of them are based on one or another version of the peaking approximation, introduced originally by Schiff . In the peaking approximation, the photon is radiated along either the initial or final electron directions, i.e. the direction of the electron is not changed while radiating, only its energy is changed. Below, we start by giving such a recipe, based on the formulas presented in this paper. What one essentially needs for a Monte Carlo simulation is an electron energy loss distribution due to real internal radiative effects. For each event one can then sample in such a distribution, both for the incoming and the outgoing electron. We next give a comparison between such a method based on the peaking approximation, with an exact numerical calculation of the radiative tail. We show to what extent the full calculation validates the approximate method for the case of elastic electron-nucleon scattering, and show that this method is realistic enough to apply it next to the calculation of the radiative tail in the case of the VCS. ### A Energy loss distribution for real internal radiative effects The details of the calculation of the real radiative corrections can be found in appendix A. It is discussed there how the real internal radiative corrections give rise to a correction factor $`e^{\delta _R}`$ to the cross section. The part of $`\delta _R`$ giving rise to the radiative tail (when differentiating $`\delta _R`$ with respect to the electron energy loss) is the first term of Eq. (A102), which contains the maximal energy of the emitted photon $`\mathrm{\Delta }E_s`$, which is defined as in Eq. (A66). The correction factor $`e^{\delta _R}`$ can be written as the product of a number of factors, of which the first one is given by : $$\left(\frac{(\mathrm{\Delta }E_s)^2}{\stackrel{~}{E}_e\stackrel{~}{E}_e^{^{}}}\right)^a,$$ (161) where $`a`$ is given by (see Eq. (A102)) : $$a=\frac{\alpha _{em}}{\pi }\left[\mathrm{ln}\left(\frac{Q^2}{m^2}\right)1\right],$$ (162) and where the tilded quantities in Eq. (161) are expressed in the c.m. system of (soft photon + recoiling proton) as explained in appendix A 5. Because in a simulation it is more straightforward to apply radiative effects in the lab, we express Eq. (161) in lab quantities, by using Eq. (A66), which yields : $$\left(\frac{(\eta \mathrm{\Delta }E_e^{})^2}{E_eE_e^{^{}}}\right)^a,$$ (163) where $`\mathrm{\Delta }E_e^{}=E_e^{el}E_e^{}`$. Introducing furthermore the quantity $`\mathrm{\Delta }E_e=\eta ^2\mathrm{\Delta }E_e^{}`$, we can write Eq. (163) as : $$\left(\frac{(\eta \mathrm{\Delta }E_e^{})^2}{E_eE_e^{^{}}}\right)^a=\left(\frac{\mathrm{\Delta }E_e\mathrm{\Delta }E_e^{}}{E_eE_e^{}}\right)^a=\left(\frac{\mathrm{\Delta }E_e}{E_e}\right)^a\left(\frac{\mathrm{\Delta }E_e^{}}{E_e^{}}\right)^a.$$ (164) The energy changes $`\mathrm{\Delta }E_e`$ ($`\mathrm{\Delta }E_e^{}`$) can be interpreted as the energy losses of the incoming (outgoing) electron due to radiation before (after) the scattering process respectively. We can then interpret the factor $`\left(\frac{\mathrm{\Delta }E_e}{E_e}\right)^a`$ as the fraction of incoming electrons which have lost an energy between 0 and $`\mathrm{\Delta }E_e`$, after being subject to real internal radiation in an equivalent radiator with thickness $`a`$. The factor $`\left(\frac{\mathrm{\Delta }E_e^{}}{E_e^{}}\right)^a`$ has a similar interpretation, but then on the outgoing electron side <sup>*</sup><sup>*</sup>*Note that when applying Eq. (164) to the radiative tail, i.e. when considering the emission of a photon whose energy is not very small compared with the electron energies, we calculate $`E_e^{el}`$ in the formula for $`\mathrm{\Delta }E_e^{}`$, using the elastic scattered energy corresponding with an initial electron which has radiated and whose energy is given by $`E_e\mathrm{\Delta }E_e`$. In the soft-photon limit this difference disappears.. Given this interpretation, if one uses a $`\mathrm{\Delta }E`$ distribution $`I_{int}(E,\mathrm{\Delta }E,a)`$, which satisfies : $$_0^{\mathrm{\Delta }E}I_{int}(E,\mathrm{\Delta }E,a)d(\mathrm{\Delta }E)=\left(\frac{\mathrm{\Delta }E}{E}\right)^a,$$ (165) then it is clear that by sampling such a distribution in a Monte Carlo simulation, the correction factor is correctly obtained. The distribution $`I_{int}`$, which has this property is given by : $$I_{int}(E,\mathrm{\Delta }E,a)=\frac{a}{\mathrm{\Delta }E}\left(\frac{\mathrm{\Delta }E}{E}\right)^a,$$ (166) and is normalized to 1 : $$_0^EI_{int}(E,\mathrm{\Delta }E,a)d(\mathrm{\Delta }E)=1.$$ (167) ### B Evaluation of the radiative tail and comparison with an exact numerical calculation for elastic electron-proton scattering Given the above distribution, a method for introducing a radiative tail due to internal radiation in a Monte Carlo simulation for elastic electron scattering suggests itself : i) For the incoming electron, sample an energy loss $`\mathrm{\Delta }E_e`$ using the distribution (166) with $`E=E_e`$ the incoming electron energy. ii) Apply elastic electron scattering using the reduced electron energy $`E_e\mathrm{\Delta }E_e`$, and if the cross section behavior is implemented in the simulation, use the elastic scattering cross section at the reduced electron energy. After the elastic scattering process, the outgoing electron has an energy $`E_e^{el}`$. iii) For the outgoing electron, sample an energy loss $`\mathrm{\Delta }E_e^{}`$ using the distribution (166) with $`E=E_e^{el}`$. The final electron energy is now $`E_e^{el}\mathrm{\Delta }E_e^{}`$. To calculate the equivalent radiator thickness $`a`$ of Eq. (162), one needs the value of $`Q^2`$, which one can in principle only calculate after the complete process has taken place. However, one can show that the above procedure reproduces the correction factor (163) with a very good accuracy already by calculating the value of $`Q^2`$ with elastic electron scattering kinematics. It is intuitively clear that the above procedure, in the case where a constant cross section is used, will reproduce the correction factor of Eq. (163). In case the actual elastic scattering cross section behavior is implemented, the cross section “walk” with the incoming electron energy is taken into account. Remark that the above procedure implies an electron energy loss both at the incoming and the outgoing electron sides. The discussed method implies, however, the assumption of a strict alignment of the bremsstrahlung photons in the direction of the radiating leptons, which is known as the (angular) peaking approximation. The strength on the other hand is found by integrating the correct angular shape in the soft photon limit, as done in appendix A 5. To test the validity of this approximate procedure, we performed a fully numerical calculation of the radiative tail for elastic electron-proton scattering. It consists of integrating over the photon phase space in the diagrams where a photon is emitted from an electron (cfr. BH diagrams of Fig. 1 (a) and (b)), as well as the diagrams where a photon is emitted from the nucleon (cfr. Born diagrams of Fig. 1 (c) and (d)). In doing so, we nowhere neglect the photon momentum $`l`$, in contrast to the calculation of appendix A 5 in the soft-photon limit, where this momentum is neglected with respect to the lepton momenta. For fixed electron kinematics, the angular phase space of the soft photon is covered by a grid with about 225000 points, chosen with increased density in the peak regions in order to keep the point-to-point change of the cross section smaller than 10 %. Attention has to be paid right in the middle of the peaks where the cross section drops very rapidly to (practically) zero within the characteristic angle $`m/E_e`$, as shown in Fig. 8. More details on this numerical integration can be found in . The result of this integration is the absolute cross section of the radiative tail, differential in the outgoing electron’s momentum and angles. It is shown by the points in Fig. 9 for $`E_e=`$ 855.0 MeV and $`\theta _e`$ = 52.18<sup>o</sup>. The energy of the outgoing electron is then determined by $`E_e^{}=E_e^{el}\mathrm{\Delta }E_e^{}`$. The points are compared with the analytical result in the soft-photon limit, obtained by differentiating the expression of Eq. (A102) for $`\delta _R`$ \- for photon emission from the electron - with respect to $`\mathrm{\Delta }E_e^{}`$. This gives a strict $`\mathrm{\Delta }E_e^1`$ behaviour, yielding the cross section $`\sigma _a\sigma _{Born}a/\mathrm{\Delta }E_e^{}`$ where the proportionality factor $`a`$ is given as in Eq. (162). The soft-photon formula gives thus a straight line when both the cross section and $`\mathrm{\Delta }E_e^{}`$ are presented on a logarithmic scale. The deviation can be seen in the lower plot of Fig. 9. From the keV-region up to about 1 MeV for $`\mathrm{\Delta }E_e^{}`$, the deviation is less then $`10^3`$ which can be taken as an upper limit for the error of the numerical integration procedure. This agreement demonstrates that the soft-photon approximation holds to very good precision in this region. For higher values of $`\mathrm{\Delta }E_e^{}`$, a raise of the photon emission cross section is observed as is expected due to the change of kinematics leading to a lower momentum transfer to the proton, and to a resulting “walk” of the cross section. We also show on the lower plot of Fig. 9 the result when both radiation from the electron and proton are considered. For better presentation, both results are normalized to the cross section $`\sigma _a`$ for soft-photon emission from the electron, as defined above. In Fig. 10, we compare for two kinematics the exact numerical calculation of the radiative tail with the approximate method of the Monte Carlo simulation as discussed above. The simulation has been investigated by running it with and without the cross section behaviour (dipole form factors assumed), and the ratio between the two versions is presented by the lines, the outer lines representing the statistcal accuracy. One notices that the increase of the radiative tail is reproduced, but somewhat overestimated compared with the exact calculation. ### C Application to virtual Compton scattering The above procedure can also be applied to VCS, as long as the angular peaking approximation is used, i.e. the electron does not change its direction while losing energy by internal real radiation. Indeed, Eq. (138) is completely similar to the elastic case, when expressing it in the c.m. system of either (soft photon + outgoing nucleon) or (soft-photon \+ outgoing photon) depending on how the $`epep\gamma `$ reaction is measured, as explained in section III B. After exponentiation, one can apply a factorization completely similar as in Eq. (164). Because under the assumption of the angular peaking approximation $`\mathrm{\Delta }E_s/\stackrel{~}{E}`$ is constant under a Lorentz transformation, we obtain the property that the shape of the distribution (166) is system independent, only its endpoint value $`E`$ changes. As a result, one can apply the distribution of Eq. (166) in the lab for VCS, but then using lab values for $`E_e`$ and $`E_e^{}`$. For VCS, one certainly can have a changing acceptance of the detection apparatus as a function of missing mass (making a “clean” cut in missing mass on the data impossible), so that generating a radiative tail in a Monte Carlo simulation with the above described method is probably the best way to implement the radiative tail correction to the data. Such a simulation was implemented for the VCS experiments already performed at MAMI and at JLab , and will be fully described in a forthcoming paper . ## V Results and discussion ### A Elastic electron-proton scattering Before showing results for VCS, we briefly discuss first the effect of the radiative corrections to elastic electron-proton scattering, in order to have a point of reference. The radiative corrections to elastic electron-proton scattering are presented in detail in appendix A. In Table I, we show for different elastic kinematics (MAMI, JLab) the numerical values of the vertex correction ($`\delta _{vertex}`$ of Eq. (A108)), the vacuum polarization correction ($`\delta _{vac}`$ of Eq. (A110)), and the real radiative correction at the electron side ($`\delta _R`$ of Eq. (A102)). We also show the $`Z`$ and $`Z^2`$ dependent corrections, $`\delta _1`$ (Eq. (A116)) and $`\delta _2^{(0)}`$ (Eq. (A119)) respectively, as derived in the recent work of . We omit here the small part in the $`Z^2`$ dependent correction which depends on the particular model for the nucleon structure (in the elastic case, the form factors), as can be found in . In Table I, we indicate the total radiative correction $`\delta _{tot}`$ as the sum of all the different contributions as in Eq. (A115). From Table I, we see that by far the largest contribution to the radiative correction comes from the large logarithm and double logarithm in $`Q^2/m^2`$ in the electron vertex correction. When evaluating the real radiative corrections for $`E_e^{el}E_e^{}=0.01E_e`$, the total effect of the radiative correction is an upwards correction of the data (for negative $`\delta _{tot}`$) of the order 20 - 25 %. In the last column of Table I (denoted by EXP), we also indicate the result when exponentiating all corrections except the vacuum polarization contribution, which - as modification of the photon propagator - is resummed as in Eq. (A114). One sees that this can lead to differences of the order of 2 %. In an elastic scattering experiment, one measures an scattered electron spectrum and has to evaluate the real radiative corrections as a function of the cut $`(E_e^{el}E_e^{})`$ which one performs in the spectrum. Dividing the measured cross section by the correction factor ($`1+\delta _{tot}`$) and plotting the result as function of $`(E_e^{el}E_e^{})`$, should then lead to a “plateau” behavior, which demonstrates the consistency of the procedure (within a certain range of the value $`(E_e^{el}E_e^{})`$ where one knows the radiative tail to sufficient accuracy). The determination of the elastic cross section for the kinematics $`E_e`$ = 705.11 MeV, $`\theta _e`$ = 42.6<sup>o</sup> is shown in Fig. 12. The upper plot shows the dE-spectrum of elastic data taken (during the beam time of the VCS experiment) at MAMI. It is compared with the simulated spectrum (dashed line). On the lower plot, the ratio of the experimental spectrum integrated up to the value $`\mathrm{\Delta }E_e^{}`$, to the simulation integrated also up to $`\mathrm{\Delta }E_e^{}`$ is shown as function of the cutoff energy $`\mathrm{\Delta }E_e^{}`$. This gives the elastic cross section, which is seen to be stable below the 1%-level over a long interval up to the cut by the acceptance of the spectrometer. The slow descent for higher $`\mathrm{\Delta }E_e^{}`$ indicates that the simulation overestimates slightly the radiative tail. ### B VCS below pion production threshold We next turn to the $`epep\gamma `$ reaction below pion threshold. It was discussed in section II, that the lowest order (in $`\alpha _{em}`$) amplitude of the $`epep\gamma `$ process at low outgoing photon energies $`\mathrm{q}^{}|\stackrel{}{q}^{^{}}|`$ is given by the BH + Born processes. The deviation from the BH + Born amplitudes grows with $`\mathrm{q}^{}`$, and can be parametrized (at low $`\mathrm{q}^{}`$) in terms of six generalized polarizabilities (GP’s) of the nucleon, which are function of $`Q^2`$. A first VCS experiment has been performed at MAMI . It consisted of measuring the $`epep\gamma `$ reaction at five values of $`\mathrm{q}^{}`$ below pion threshold, ranging from $`\mathrm{q}^{}`$ = 33 MeV/c to $`\mathrm{q}^{}`$ = 111.5 MeV/c. At the lowest value $`\mathrm{q}^{}`$ = 33 MeV/c, where the polarizability effect is negligeably small, the measurement serves as a check of the Low Energy Theorem (LET). The measured deviation as function of $`\mathrm{q}^{}`$ can then be interpreted as the effect of the GP’s. It is clear that both to test the LET as well as to extract the GP’s from the measured deviation with respect to the BH + Born result (which is expected to be of the order 10 - 20 % at the highest $`\mathrm{q}^{}`$ value), it is a prerequisite to know very accurately how the result is modified due to radiative corrections. In Fig. 13, we first show the differential cross section for MAMI kinematics at a low value $`\mathrm{q}^{}`$ = 33 MeV/c, as function of the c.m. angle of the emitted real photon with respect to the direction of the virtual photon. One sees from Fig. 13 that the virtual radiative corrections reduce the BH + Born result in these kinematics by about 16 % (or when applied to data, increase the uncorrected data by 16 %). The real radiative corrections have to be estimated as function of the cut which one performs in the missing mass spectrum. The VCS experiments below pion threshold measure the $`epep\gamma `$ reaction by detecting the outgoing electron and proton, and reconstruct the missing mass $`M_{m2}`$ as defined in Eq. (144). In Fig. 13, the real radiative corrections are shown for a value of $`\mathrm{\Delta }E_s`$ = 10 MeV, where the soft-photon energy $`\mathrm{\Delta }E_s`$ is determined from the cut in the missing mass according to Eq. (148). For the small value $`\mathrm{q}^{}`$ = 33 MeV/c, the real radiative correction depends only very little on the angle $`\theta _{\gamma \gamma }`$ (through the last terms on the rhs of Eqs. (145,146)). For $`\mathrm{\Delta }E_s`$ = 10 MeV, the real radiative correction $`\delta _R`$ is given by $`\delta _R`$ \- 0.025, which corresponds with increasing the uncorrected data by about 2.5 %. For $`\mathrm{\Delta }E_s`$ = 20 MeV, $`\delta _R`$ \+ 0.02 (reducing the uncorrected data by about 2 %), and for $`\mathrm{\Delta }E_s`$ = 30 MeV, $`\delta _R`$ \+ 0.045 (reducing the uncorrected data by about 4.5 %). To determine the $`epep\gamma `$ cross section from the measured missing mass spectra, one has to perform a consistency check by plotting the experimentally measured (uncorrected) cross section divided by the radiative correction factor as function of the cut in the missing mass spectrum. In this way, one has to find a “plateau” behavior, as was demonstrated before for elastic data. This consistency check was also performed on the VCS data measured at MAMI , and will be shown in a forthcoming publication . In Fig. 14, we show the the differential cross section for MAMI kinematics at the highest measured outgoing photon energy : $`\mathrm{q}^{}`$ = 111.5 MeV/c. The virtual radiative corrections are mainly $`\mathrm{q}^{}`$ independent (for these rather small values) and lead thus also here to a reduction of the BH + Born result by about 16 %. The real radiative corrections are again shown for $`\mathrm{\Delta }E_s`$ = 10 MeV, and exhibit a slight angular dependence. These corrections were applied to the data from the unpolarized MAMI experiment of . From the deviation of the radiatively corrected data and the BH + Born result, two combinations of GP’s have been extracted at $`Q^2`$ 0.33 GeV<sup>2</sup> in . An experiment below pion production threshold to measure the GP’s at higher $`Q^2`$ has also been measured at JLab and is under analysis at the time of writing. In Fig. 15, we show how the BH + Born cross section is modified due to the virtual radiative corrections. It is seen that for the JLab kinematics of Fig. 15, the BH + Born result is reduced at the backward angles by about 20 % due to the virtual radiative corrections. The unpolarized VCS cross section below pion threshold provides three independent structure functions (when varying the value of $`\epsilon `$ in the experiment), which allows to extract three of the six (lowest order) generalized nucleon polarizabilities. To extract the three remaining nucleon polarizabilities, one has to resort to double polarization observables as discussed in . In particular, double polarization observables with polarized electron beam and with a polarized target (along either of the three axes), or alternatively by measuring the recoil nucleon polarization, provide three new observables to extract the three additional nucleon response functions . In Fig. 16, we show the double polarization asymmetries for MAMI kinematics, by measuring the recoil polarization components along the $`z`$-direction (virtual photon direction) or along the $`x`$-direction (perpendicular to the virtual photon but parallel to the scattering plane). One aims to extract the polarizability effect in these observables from the deviation of the measured asymmetry and the BH + Born result (see e.g. for an estimate of this effect within a model calculation). Therefore, it is important to know how much the BH + Born result is affected by the radiative corrections before extracting the polarizability effect. It is seen in Fig. 16 that the effect of the radiative corrections on the double polarization asymmetries nearly drops out in the ratio (much less than 1 % change of the asymmetries). At the low values of the outgoing photon energy $`\mathrm{q}^{}`$ (e.g. $`\mathrm{q}^{}`$ 33 MeV/c) where the polarizability effect is very small, these asymmetries are also hardly affected by radiative corrections. Therefore, these asymmetries can also provide an independent check of the LET. An experiment to measure the VCS double polarization observables by measuring the recoil nucleon polarization is planned at MAMI in the near future . ### C Deeply virtual Compton scattering Besides the low energy region, the VCS process is also studied in the Bjorken regime, where $`Q^2`$ and $`\nu =p.q/M_N`$ are large, with $`x_B=Q^2/(2M_N\nu )`$ fixed. In this kinematical region, the process is refered to as deeply virtual Compton scattering (DVCS). In the Bjorken regime, the DVCS amplitude factorizes into a perturbatively calculable hard scattering amplitude, and into a non-perturbative part at the proton side, expressed in terms of so-called skewed parton distributions (SPD’s) which generalize the ordinary parton distributions. These SPD’s are new nucleon structure observables which one aims to extract by measuring e.g. the exclusive $`epep\gamma `$ reaction in the Bjorken regime. Similarly as was seen before in the threshold region, the $`epep\gamma `$ reaction can have an important contribution from the BH process, besides the DVCS process of actual interest. However, the BH and DVCS contributions behave differently as function of the lepton beam energy, as studied in Refs. . In particular, at the lower beam energies, such as e.g. available at JLab, the BH process dominates in the forward direction over the DVCS process. In this region, the DVCS process becomes only measurable due to its interference with the BH process. In order to extract the DVCS process (and the nucleon structure information) from its interference with the BH, it is therefore important to have good knowledge of how the radiative corrections modify the BH amplitude. In Fig. 17, we show the $`epep\gamma `$ cross section in kinematics accessible at JLab, where such an experiment is planned . The DVCS cross section is calculated by using the ansatz for the SPD’s of . It is seen from Fig. 17, that the BH indeed dominates over the DVCS cross section in these kinematics, and that the DVCS cross section gets enhanced due to its interference with the BH. One furthermore sees that the virtual radiative corrections reduce the BH + DVCS cross section by about 23% in these kinematics. This is mainly due to the reduction of the BH process when including virtual radiative corrections. The real radiative corrections are shown in Fig. 17 for a value $`\mathrm{\Delta }E_s`$ = 0.1 GeV, which corresponds with a cut in the recoiling hadronic missing mass spectrum (defined in Eq. (139)) of $`M_{m1}^2`$ \- $`M_N^2`$ 0.21 GeV<sup>2</sup>. In Ref. , it was suggested that an exploratory study of the DVCS process might be possible by studying the $`epep\gamma `$ reaction with a polarized electron beam. The electron single spin asymmetry (SSA) does not vanish out of plane and is only due to the interference of the BH amplitude and the imaginary part of the DVCS amplitude (i.e. the BH amplitude does not lead to a SSA, because it is purely real). Therefore, one expects this SSA to be less sensitive to radiative corrections on the BH amplitude. However, as the BH amplitude enters the SSA linearly in the numerator, but quadratically in the denominator (as in the unpolarized cross section), one might wonder what is the residual effect of the radiative corrections on this observable. In Fig. 18, we show the SSA for DVCS at JLab. One sees that the SSA gets only slightly reduced due to the radiative corrections. The reduction of the SSA amounts to maximum 5% of its value around 5<sup>o</sup>, where the asymmetry reaches its maximal value. Therefore, the SSA shows to be a rather “clean” observable for extracting the DVCS amplitude in a region where the BH process dominates. Its measurement is also envisaged at JLab in the near future . ## VI Conclusions We studied in this work the first order QED radiative corrections to the $`epep\gamma `$ reaction. The one-loop virtual radiative corrections have been evaluated by a combined analytical-numerical method. Several tests were shown to cross-check the numerical method used. Furthermore, it was shown how all IR divergences cancel when adding the soft-photon emission processes. A fully numerical method was presented for the photon emission processes where the photon energy is not very small compared with the electron energies, which makes up the radiative tail. Besides, we have also presented an approximate calculation of the radiative tail, which was shown to be realistic enough for use in a Monte Carlo simulation. We compared our results first to elastic electron-proton scattering. Subsequently, the results for the radiative corrections to the $`epep\gamma `$ reaction were shown both below pion threshold and in the deeply virtual Compton scattering regime. Below pion threshold, our calculations were applied to the first dedicated VCS experiment at MAMI, and show that the effect of the radiative corrections results in an enhancement of the uncorrected data by about 20 % (or an equivalent reduction of the theory). VCS double polarization asymmetries where shown to be insensitive to radiative corrections. For the DVCS, we calculated radiative corrections for JLab kinematics and found the virtual radiative corrections to lead to an enhancement of the data by about 23 %. The single spin asymmetry was shown to be only slightly reduced by radiative corrections. Although we focussed here on the kinematical regimes of ongoing or planned experiments, the present work can also serve as a tool in the analysis of future VCS experiments. ## acknowledgments As this work grew over a period of a couple of years, we enjoyed discussions with many colleagues about the subject of this work, in particular with N. d’Hose and P. Vernin, about the application of radiative corrections to elastic and VCS data. We thank M. Distler and H. Merkel for their help in the numerical implementation of the radiative tail calculation. Furthermore, we would like to thank P.A.M. Guichon for many discussions, which were at the origin of this work. We also want to acknowledge in particular very useful discussions and correspondence with L. Maximon and are grateful for his continued interest in this work. This work was supported in part by the French Commissariat à l’Energie Atomique (CEA), by the EU/TMR contract ERB FMRX-CT96-0008, by the Deutsche Forschungsgemeinschaft (SFB 443), by the French CNRS/IN2P3, and by the Fund for Scientific Research - Flanders (Belgium). ## A Radiative corrections to elastic lepton-nucleon scattering using the dimensional regularization method for both UV and IR divergences. In this appendix, we provide the reader with some details of the derivation of the radiative corrections to elastic lepton scattering at one-loop level. In our derivation, we use the dimensional regularization procedure to regularize both ultraviolet and infrared divergences. After a short introduction of the renormalization method, we calculate subsequently the vertex diagram at the lepton side (Fig. 19(a)), the lepton self-energy diagram (Fig. 19(b)), the vacuum polarization diagram (Fig. 19(c)), and give an analytical result, without approximations, for the soft photon emission at the lepton side (Fig. 19 (d) and (e)). We compare our results with other derivations found in the literature. At the end we collect the results to correct the elastic lepton-nucleon scattering cross sections and discuss the role of the radiative corrections at the proton side and the two-photon exchange corrections by referring to the recent work of Ref. . In this appendix, we use the same notations as explained in section II. ### 1 Renormalization method In calculating QED radiative corrections in this work, we are using the BPHZ renormalization method (as explained e.g. in Ref.), which consists of replacing in the unrenormalized Lagrangian all bare quantities by renormalized ones. For QED, the bare Lagrangian is given by (we are using the conventions of Bjorken and Drell in this work) $$_B=\overline{\mathrm{\Psi }}_B\left(i\gamma ^\mu _\mu m_B\right)\mathrm{\Psi }_B\frac{1}{4}F_{B\mu \nu }F_{B}^{}{}_{}{}^{\mu \nu }e_B\overline{\mathrm{\Psi }}_B\gamma ^\mu \mathrm{\Psi }_BA_{B\mu },$$ (A1) where the bare field tensor $`F_B^{\mu \nu }`$ is given by $$F_B^{\mu \nu }=^\mu A_B^\nu ^\nu A_B^\mu .$$ (A2) The renormalization of the theory amounts in redefining the bare quantities in terms of renormalized (i.e. physical observable) ones : $`\mathrm{\Psi }_B=Z_2^{1/2}\mathrm{\Psi },A_{B}^{}{}_{}{}^{\mu }=Z_3^{1/2}A^\mu ,`$ (A3) $`m_B=Z_mm,e_B=Z_ge.`$ (A4) In Eq. (A4), the renormalized finite quantities are $`\mathrm{\Psi },A^\mu ,m`$ and $`e`$. A theory in which all divergences can be absorbed into renormalization constants such as $`Z_2,Z_3,Z_m`$ and $`Z_g`$ in Eq. (A4), is called multiplicatively renormalizable. This procedure leads to a decomposition of the QED Lagrangian of Eq. (A1) into $$_B=_R+_{CT},$$ (A5) where $`_R`$ represents the renormalized Lagrangian in terms of the physical (finite) quantities $$_R=\overline{\mathrm{\Psi }}\left(i\gamma ^\mu _\mu m\right)\mathrm{\Psi }\frac{1}{4}F_{\mu \nu }F^{\mu \nu }e\overline{\mathrm{\Psi }}\gamma ^\mu \mathrm{\Psi }A_\mu ,$$ (A6) and where $`_{CT}`$ is called the counterterm Lagrangian $$_{CT}=(Z_21)\overline{\mathrm{\Psi }}i\gamma ^\mu _\mu \mathrm{\Psi }(Z_2Z_m1)\overline{\mathrm{\Psi }}m\mathrm{\Psi }(Z_31)\frac{1}{4}F_{\mu \nu }F^{\mu \nu }(Z_11)e\overline{\mathrm{\Psi }}\gamma ^\mu \mathrm{\Psi }A_\mu .$$ (A7) In Eq. (A7), the vertex renormalization constant $`Z_1`$ is defined as $`Z_1=Z_gZ_2Z_3^{1/2}`$. For a renormalizable theory such as QED, all divergences obtained by calculating loop diagrams with the renormalized Lagrangian $`_R`$ are cancelled by the corresponding contributions in the counterterm Lagrangian $`_{CT}`$. It will be shown below how the QED renormalization constants are calculated to order $`O(e^2)`$ by calculating the vertex diagram, the lepton self-energy diagram and the photon polarization diagram at the one-loop level. As QED is a gauge invariant theory, we will simplify all calculations in this work by using the Feynman gauge. ### 2 Vertex diagram The on-shell photon-lepton-lepton vertex is represented by $$M_v^\mu =\overline{u}(k^{^{}},h^{})\left[ie\mathrm{\Lambda }^\mu (k^{^{}},k)\right]u(k,h),$$ (A8) and the on-shell vertex of Eq. (A8) can be parametrized as $$\overline{u}(k^{^{}},h^{})\mathrm{\Lambda }^\mu (k^{^{}},k)u(k,h)=\overline{u}(k^{^{}},h^{})\left[\left(1+F(Q^2)\right)\gamma ^\mu G(Q^2)i\sigma ^{\mu \nu }\frac{q_\nu }{2m}\right]u(k,h),$$ (A9) where $`q=kk^{^{}}`$. To order $`O(e^2)`$ , the vertex $`\mathrm{\Lambda }^\mu `$ (corresponding with Fig. 19(a)) is given by $$\mathrm{\Lambda }^\mu (k^{^{}},k)=\gamma ^\mu ie^2\mu ^{4D}\frac{d^Dl}{\left(2\pi \right)^D}\frac{\gamma ^\alpha \left(\mathit{}^{^{}}+\mathit{}+m\right)\gamma ^\mu \left(\mathit{}+\mathit{}+m\right)\gamma _\alpha }{\left[l^2\right][l^2+2l.k^{^{}}][l^2+2l.k]}+O\left(e^4\right),$$ (A10) where a mass scale $`\mu `$ (renormalization scale) has to be introduced when passing to $`D4`$ dimensions in order to keep the coupling constant dimensionless. It is immediately seen by power counting that in four dimensions ($`D=4`$), the one-loop integral in Eq. (A10) contains an ultraviolet ($`l\mathrm{}`$) logarithmic divergence and an infrared ($`l0`$) logarithmic divergence. To subtract the divergent parts (by the corresponding counterterms) of expressions such as Eq. (A10), one has to regularize them first. We follow in this work the dimensional regularization procedure to regularize both ultraviolet and infrared divergences. The dimensional regularization method amounts in calculating loop diagrams in $`D`$ dimensions. Physical observables are obtained by letting $`D4`$ at the end. To obtain an integral which is ultraviolet convergent, one has to take $`D<4`$, or $`ϵ_{UV}2D/2>0`$ in expressions such as Eq. (A10). To obtain an integral which is infrared convergent, one has to take $`D>4`$, or $`ϵ_{IR}2D/2<0`$. The two different limits show that care has to be taken with the limit $`D4`$, which means that the parts in Eq. (A10) that are infrared divergent and the parts that are ultraviolet divergent have to be separated and in the corresponding terms, two different limits have to be taken when one approaches $`D=4`$. Although the dimensional regularization scheme has been applied originally to ultraviolet divergent expressions as it respects the symmetries of the theory (in particular the gauge symmetry for a gauge theory), it has also been applied in a few works to regularize infrared divergences . When working out the integral in Eq. (A10), one obtains after some algebra the following expressions for $`F(Q^2)`$ and $`G(Q^2)`$ to order $`O(e^2)`$ : $`F(Q^2)={\displaystyle \frac{e^2}{\left(4\pi \right)^2}}`$ $`\{[{\displaystyle \frac{1}{\epsilon _{UV}}}\gamma _E+\mathrm{ln}\left({\displaystyle \frac{4\pi \mu ^2}{m^2}}\right)]+[{\displaystyle \frac{1}{\epsilon _{IR}}}\gamma _E+\mathrm{ln}\left({\displaystyle \frac{4\pi \mu ^2}{m^2}}\right)]{\displaystyle \frac{v^2+1}{v}}\mathrm{ln}\left({\displaystyle \frac{v+1}{v1}}\right)`$ (A13) $`+{\displaystyle \frac{v^2+1}{2v}}\mathrm{ln}\left({\displaystyle \frac{v+1}{v1}}\right)\mathrm{ln}\left({\displaystyle \frac{v^21}{4v^2}}\right)+{\displaystyle \frac{2v^2+1}{v}}\mathrm{ln}\left({\displaystyle \frac{v+1}{v1}}\right)`$ $`+{\displaystyle \frac{v^2+1}{v}}[Sp\left({\displaystyle \frac{v+1}{2v}}\right)Sp\left({\displaystyle \frac{v1}{2v}}\right)]\},`$ and $$G(Q^2)=\frac{e^2}{\left(4\pi \right)^2}\frac{v^21}{v}\mathrm{ln}\left(\frac{v+1}{v1}\right),$$ (A14) where $`v`$ is given by $$v^2\mathrm{\hspace{0.33em}1}+\frac{4m^2}{Q^2},$$ (A15) with $`Q^2`$ = - $`q^2`$ $`>`$ 0. In Eq. (A13), $`\gamma _E`$ represents the Euler constant, and the Spence (or dilogarithmic) function is defined by $$Sp(x)\underset{0}{\overset{x}{}}𝑑t\frac{\mathrm{ln}(1t)}{t}.$$ (A16) From Eq. (A14), the one-loop radiative correction to the electron magnetic moment follows as $$\mu =\frac{e}{2m}\left(1+G(Q^2=\mathrm{\hspace{0.17em}0})\right)=\frac{e}{2m}\left(1+\frac{\alpha _{em}}{2\pi }\right),$$ (A17) which is the result first obtained by Schwinger . To remove the UV divergence from the vertex correction Eq. (A13), one has to determine the vertex renormalization constant $`Z_1`$ of Eq. (A7). $`Z_1`$ is determined by requiring that the total vertex $$\stackrel{~}{\mathrm{\Lambda }}^\mu =\mathrm{\Lambda }^\mu +(Z_1\mathrm{\hspace{0.17em}1})\gamma ^\mu ,$$ (A18) defines the physical electron charge at $`Q^2=0`$, i.e. $`Z_1=\mathrm{\hspace{0.17em}1}F(Q^2=\mathrm{\hspace{0.17em}0})`$ (A19) $`=1{\displaystyle \frac{e^2}{\left(4\pi \right)^2}}\left\{\left[{\displaystyle \frac{1}{\epsilon _{UV}}}\gamma _E+\mathrm{ln}\left({\displaystyle \frac{4\pi \mu ^2}{m^2}}\right)\right]+\mathrm{\hspace{0.17em}2}\left[{\displaystyle \frac{1}{\epsilon _{IR}}}\gamma _E+\mathrm{ln}\left({\displaystyle \frac{4\pi \mu ^2}{m^2}}\right)\right]+\mathrm{\hspace{0.17em}4}\right\}+O(e^4).`$ (A20) It is seen that the vertex renormalization constant $`Z_1`$ contains besides the UV divergence also an IR divergence. The renormalized vertex of Eq. (A18), is determined by the vertex correction function $`F(Q^2)F(Q^2=0)`$ which is given to first order in $`\alpha _{em}`$ (where $`\alpha _{em}=e^2/4\pi `$) by the expression $`F(Q^2)F(Q^2=0)=`$ $`{\displaystyle \frac{\alpha _{em}}{2\pi }}\{[{\displaystyle \frac{1}{\epsilon _{IR}}}\gamma _E+\mathrm{ln}\left({\displaystyle \frac{4\pi \mu ^2}{m^2}}\right)].[{\displaystyle \frac{v^2+1}{2v}}\mathrm{ln}\left({\displaystyle \frac{v+1}{v1}}\right)\mathrm{\hspace{0.17em}1}]`$ (A23) $`+{\displaystyle \frac{v^2+1}{4v}}\mathrm{ln}\left({\displaystyle \frac{v+1}{v1}}\right)\mathrm{ln}\left({\displaystyle \frac{v^21}{4v^2}}\right)+{\displaystyle \frac{2v^2+1}{2v}}\mathrm{ln}\left({\displaystyle \frac{v+1}{v1}}\right)2`$ $`+{\displaystyle \frac{v^2+1}{2v}}[Sp\left({\displaystyle \frac{v+1}{2v}}\right)Sp\left({\displaystyle \frac{v1}{2v}}\right)]\},`$ The expression for the vertex correction function $`F(Q^2)F(Q^2=0)`$, which was calculated here using the dimensional regularization method for both the UV and IR divergences, agrees with the ones derived in many textbooks (see e.g. Eq. (47.52) of Ref. where a full derivation is given). The correspondence with the calculations which use a finite photon mass ($`\lambda `$) as IR regulator is found to be $$\frac{1}{ϵ_{IR}}\gamma _E+\mathrm{ln}\left(\frac{4\pi \mu ^2}{m^2}\right)\mathrm{ln}\frac{\lambda ^2}{m^2}.$$ (A24) In the ultrarelativistic limit ($`Q^2>>m^2`$), the vertex correction $`F(Q^2)F(Q^2=0)`$ can be found from Eq. (A23) to be given by $`F(Q^2)F(Q^2=0)\stackrel{Q^2>>m^2}{}{\displaystyle \frac{\alpha _{em}}{2\pi }}`$ $`\{[{\displaystyle \frac{1}{\epsilon _{IR}}}\gamma _E+\mathrm{ln}\left({\displaystyle \frac{4\pi \mu ^2}{m^2}}\right)].[\mathrm{ln}\left({\displaystyle \frac{Q^2}{m^2}}\right)1]`$ (A26) $`+({\displaystyle \frac{3}{2}}\mathrm{ln}\left({\displaystyle \frac{Q^2}{m^2}}\right)2)+({\displaystyle \frac{1}{2}}\mathrm{ln}^2\left({\displaystyle \frac{Q^2}{m^2}}\right)+{\displaystyle \frac{\pi ^2}{6}})\}.`$ It is seen from Eq. (A26), that the finite part of the vertex correction at high $`Q^2`$ is dominated by a quadratic logarithmic term. ### 3 Lepton self-energy diagram The free lepton propagator (for a lepton with four-momentum $`k`$) $$S^o\left(k\right)=\frac{\mathit{}+m}{k^2m^2+iϵ},$$ (A27) is modified through the lepton self-energy $`\mathrm{\Sigma }\left(k\right)`$, to the full lepton propagator $$S\left(k\right)=S^o\left(k\right)+S^o\left(k\right)\mathrm{\Sigma }\left(k\right)S\left(k\right).$$ (A28) To first order $`O(e^2)`$, the lepton self-energy (corresponding with Fig.19(b)) is given by $$i\mathrm{\Sigma }\left(k\right)=e^2\mu ^{4D}\frac{d^Dl}{\left(2\pi \right)^D}\frac{\gamma ^\alpha \left(\mathit{}+\mathit{}+m\right)\gamma _\alpha }{\left[l^2\right]\left[(k+l)^2m^2\right]}.$$ (A29) By power counting, it is seen that the integral of Eq. (A29) contains a linear UV divergence but is IR finite in the limit $`D4`$. The integral of Eq. (A29) can be worked out and yields $`\mathrm{\Sigma }\left(k\right)={\displaystyle \frac{e^2}{\left(4\pi \right)^2}}`$ $`\{[{\displaystyle \frac{1}{\epsilon _{UV}}}\gamma _E+\mathrm{ln}\left({\displaystyle \frac{4\pi \mu ^2}{m^2}}\right)](\mathit{}4m)`$ (A32) $`+\mathit{}\left[1+{\displaystyle \frac{1}{\stackrel{~}{k}^2}}+{\displaystyle \frac{1+\stackrel{~}{k}^2}{\left(\stackrel{~}{k}^2\right)^2}}(1\stackrel{~}{k}^2)\mathrm{ln}\left(1\stackrel{~}{k}^2\right)\right]`$ $`+\mathrm{\hspace{0.17em}2}m[3{\displaystyle \frac{2}{\stackrel{~}{k}^2}}(1\stackrel{~}{k}^2)\mathrm{ln}(1\stackrel{~}{k}^2)]\},`$ where $`\stackrel{~}{k}^2=k^2/m^2`$. To remove the UV divergence from the self-energy Eq. (LABEL:eq:a\_self4), one has to determine the renormalization constants $`Z_2`$ and $`Z_m`$ from Eq. (A7). This counterterm contribution leads to the renormalized self-energy $$\stackrel{~}{\mathrm{\Sigma }}\left(k\right)=\mathrm{\Sigma }\left(k\right)(Z_21)\mathit{}+(Z_2Z_m1)m.$$ (A34) Inserting Eq. (A34) into Eq. (A28) and developing $`\mathrm{\Sigma }\left(k\right)`$ as a Taylor series expansion around $`\mathit{}=m`$ yields for inverse of the total lepton propagator $`S^1`$ $`=`$ $`(\mathit{}m)\left[1{\displaystyle \frac{d\mathrm{\Sigma }}{d\mathit{}}}|_{\mathit{}=m}+(Z_21)\right]+\left[(1Z_m)Z_2m\mathrm{\Sigma }(\mathit{}=m)\right]`$ (A36) $`+O\left((\mathit{}m)^2\right).`$ Requiring that the total propagator $`S`$ has a pole at $`\mathit{}=m`$ with residue 1, determines the renormalization constants $`Z_2`$ and $`Z_m`$ as $`Z_2=\mathrm{\hspace{0.33em}1}+{\displaystyle \frac{d\mathrm{\Sigma }}{d\mathit{}}}|_{\mathit{}=m},`$ (A37) $`(1Z_m)Z_2m=\mathrm{\Sigma }(\mathit{}=m).`$ (A38) Using the first order expression of Eq. (LABEL:eq:a\_self4) for the lepton self-energy, yields $`Z_2=\mathrm{\hspace{0.33em}1}{\displaystyle \frac{e^2}{\left(4\pi \right)^2}}`$ $`\{[{\displaystyle \frac{1}{\epsilon _{UV}}}\gamma _E+\mathrm{ln}\left({\displaystyle \frac{4\pi \mu ^2}{m^2}}\right)]`$ (A40) $`+\mathrm{\hspace{0.33em}2}[{\displaystyle \frac{1}{\epsilon _{IR}}}\gamma _E+\mathrm{ln}\left({\displaystyle \frac{4\pi \mu ^2}{m^2}}\right)]+\mathrm{\hspace{0.33em}4}\}+O(e^4),`$ $`Z_2Z_m=\mathrm{\hspace{0.33em}1}{\displaystyle \frac{e^2}{\left(4\pi \right)^2}}`$ $`\{4[{\displaystyle \frac{1}{\epsilon _{UV}}}\gamma _E+\mathrm{ln}\left({\displaystyle \frac{4\pi \mu ^2}{m^2}}\right)]`$ (A42) $`+\mathrm{\hspace{0.33em}2}[{\displaystyle \frac{1}{\epsilon _{IR}}}\gamma _E+\mathrm{ln}\left({\displaystyle \frac{4\pi \mu ^2}{m^2}}\right)]+\mathrm{\hspace{0.33em}8}\}+O(e^4).`$ Remark that although the unrenormalized lepton self-energy $`\mathrm{\Sigma }(k)`$ of Eq. (A29) is IR finite, the lepton field renormalization constant $`Z_2`$ contains an infrared divergence for the derivative of $`\mathrm{\Sigma }`$ that appears in its definition (see Eq. (A38)). Furthermore, a comparison of the first order expressions for the lepton field renormalization constant $`Z_2`$ (Eq. (A40)) with the vertex renormalization constant $`Z_1`$ (Eq. (A19)) shows that they are the same (It is known as a Ward identity and can be shown to hold to all orders as a consequence of the gauge invariance of QED). Finally, using the expressions of Eqs.(A40),(A42), the renormalized lepton self-energy to first order in $`\alpha _{em}`$ is given by $`\stackrel{~}{\mathrm{\Sigma }}\left(k\right)={\displaystyle \frac{\alpha _{em}}{4\pi }}`$ $`\{\mathit{}[\mathrm{\hspace{0.33em}2}({\displaystyle \frac{1}{\epsilon _{IR}}}\gamma _E+\mathrm{ln}\left({\displaystyle \frac{4\pi \mu ^2}{m^2}}\right))3+{\displaystyle \frac{1}{\stackrel{~}{k}^2}}+{\displaystyle \frac{(1\stackrel{~}{k}^4)}{\stackrel{~}{k}^4}}\mathrm{ln}(1\stackrel{~}{k}^2)]`$ (A44) $`m[\mathrm{\hspace{0.33em}2}({\displaystyle \frac{1}{\epsilon _{IR}}}\gamma _E+\mathrm{ln}\left({\displaystyle \frac{4\pi \mu ^2}{m^2}}\right))2+{\displaystyle \frac{4}{\stackrel{~}{k}^2}}(1\stackrel{~}{k}^2)\mathrm{ln}(1\stackrel{~}{k}^2)]\}.`$ It is seen from Eq. (A44) that for an on-shell lepton ($`\mathit{}=m`$), the renormalized lepton self-energy $`\stackrel{~}{\mathrm{\Sigma }}`$ is exactly zero. Consequently, this correction has only to be applied for internal lepton lines. ### 4 Vacuum polarization diagram Starting form the free propagator of a photon with four-momentum $`q`$ (as stated before, we give all expressions in the Feynman gauge) $$D_o^{\mu \nu }(q)=\frac{g^{\mu \nu }}{q^2},$$ (A45) the full photon propagator can be written as $$D^{\mu \nu }(q)=D_o^{\mu \nu }(q)+D^{\mu \kappa }(q)\mathrm{\Pi }_{\kappa \lambda }(q)D_o^{\lambda \nu }(q),$$ (A46) where $`\mathrm{\Pi }_{\kappa \lambda }(q)`$ represents the vacuum polarization correction. To order $`O(e^2)`$, the vacuum polarization (corresponding with Fig.19(c)) due to $`l^+l^{}`$ loops (with lepton $`l=e,\mu ,\tau `$) is given by $$i\mathrm{\Pi }^{\mu \nu }(q)=e^2\mu ^{4D}\frac{d^Dl}{\left(2\pi \right)^D}\frac{Tr\left\{\gamma ^\mu \left(\mathit{}+\mathit{}+m\right)\gamma ^\nu \left(\mathit{}+m\right)\right\}}{\left[\left(l+q\right)^2m^2\right]\left[l^2m^2\right]}+O\left(e^4\right).$$ (A47) The gauge invariance of QED leads to the relation $`q^\kappa q^\lambda \mathrm{\Pi }_{\kappa \lambda }(q)=0`$ (Ward-Takahashi identity). Consequently, the vacuum polarization correction can be written as $$\mathrm{\Pi }_{\kappa \lambda }(q)=\left(g_{\kappa \lambda }q^2+q_\kappa q_\lambda \right)\mathrm{\Pi }\left(q^2\right),$$ (A48) where the function $`\mathrm{\Pi }\left(q^2\right)`$ is IR convergent and contains only a logarithmic UV divergence as can be seen from Eq. (A47). Using Eq. (A48), the self-consistent relation for the full photon propagator (Eq. (A46)) yields $$D^{\mu \nu }(q)=\frac{g^{\mu \nu }}{q^2\left(1\mathrm{\Pi }(q^2)\right)}+\mathrm{t}erminq^\mu q^\nu ,$$ (A49) where we don’t have to specify the term in $`q^\mu q^\nu `$, as the photon propagator will be contracted with conserved currents on both sides so that this term in $`q^\mu q^\nu `$ will not contribute to physical observables. Evaluating the one-loop integral of Eq. (A47), one obtains $$\mathrm{\Pi }(Q^2)=\frac{e^2}{(4\pi )^2}\frac{4}{3}\left[\frac{1}{\epsilon _{UV}}\gamma _E+\mathrm{ln}\left(\frac{4\pi \mu ^2}{m^2}\right)\left(v^2\frac{8}{3}\right)+v\frac{\left(v^23\right)}{2}\mathrm{ln}\left(\frac{v+1}{v1}\right)\right],$$ (A50) where $`v`$ is given by Eq. (A15). The UV divergent term in Eq. (A50) is removed by adding the counterterm in $`Z_3`$ of Eq. (A7). This leads to the renormalized photon propagator $$\stackrel{~}{D}^{\mu \nu }(q)=\frac{g^{\mu \nu }}{q^2\left(1\stackrel{~}{\mathrm{\Pi }}(q^2)\right)}+\mathrm{t}erminq^\mu q^\nu ,$$ (A51) where the renormalized photon polarization $`\stackrel{~}{\mathrm{\Pi }}`$ is given by $$\stackrel{~}{\mathrm{\Pi }}(Q^2)=\mathrm{\Pi }(Q^2)\left(Z_31\right).$$ (A52) Requiring that the renormalized photon propagator (Eq. (A49)) has a pole at $`q^2=0`$ with residue 1, determines the renormalization constant $`Z_3`$ : $$Z_3=\mathrm{\hspace{0.33em}1}+\mathrm{\Pi }\left(q^2=0\right).$$ (A53) Consequently, the renormalized finite photon polarization is found from Eqs.(A50) and (A52) to be given by $$\stackrel{~}{\mathrm{\Pi }}(Q^2)=\frac{\alpha _{em}}{\pi }\frac{1}{3}\left[\left(v^2\frac{8}{3}\right)+v\frac{\left(3v^2\right)}{2}\mathrm{ln}\left(\frac{v+1}{v1}\right)\right],$$ (A54) which agrees with the result derived in Ref.. ### 5 Soft photon emission contributions The calculation of the one-loop vertex correction of Eq. (A10) was seen to be both UV and IR divergent. The ultraviolet divergence was removed by renormalizing the fields and parameters of the theory. The remaining infrared divergences are cancelled at the cross section level by the soft bremsstrahlung contributions . In this bremsstralung process (see Figs.19 (d) and (e)), an electron is accompanied by the emission of a soft photon of maximal energy $`\mathrm{\Delta }E_s`$ (which is related to the detector resolution and is therefore much smaller than the electron energy which radiates this soft photon). To first order in $`\alpha _{em}`$ (relative to the Born cross section) the bremsstrahlung cross section amounts to calculate a phase space integral of the form : $`d\sigma {\displaystyle \frac{d^3\stackrel{}{k}_e^{^{}}}{\left(2\pi \right)^3\mathrm{\hspace{0.17em}2}E_e^{}}}{\displaystyle \frac{d^3\stackrel{}{p}_N^{^{}}}{\left(2\pi \right)^3\mathrm{\hspace{0.17em}2}E_N^{}}}{\displaystyle \frac{d^3\stackrel{}{l}}{\left(2\pi \right)^3\mathrm{\hspace{0.17em}2}\mathrm{l}}}`$ $`(2\pi )^4\delta ^4(k+pk^{}p^{}l)`$ (A56) $`\times |M_{BORN}|^2(e^2)[{\displaystyle \frac{k_\mu ^{^{}}}{k^{^{}}.l}}{\displaystyle \frac{k_\mu }{k.l}}].[{\displaystyle \frac{k^\mu }{k^{^{}}.l}}{\displaystyle \frac{k^\mu }{k.l}}],`$ where $`\mathrm{l}|\stackrel{}{l}|`$ denotes the soft photon energy, and where $`M_{BORN}`$ denotes the Born amplitude for elastic lepton-nucleon scattering. In Eq. (A56), terms in the soft photon momentum were neglected compared with the electron momenta $`k`$ and $`k^{^{}}`$, except in the denominators of the lepton propagators where they matter. If one performs an experiment where the outgoing electron is detected, and where the recoiling proton remains undetected (i.e. if one measures a single arm electron spectrum), one eliminates in Eq. (A56) the integral over $`\stackrel{}{p}_N^{^{}}`$ with the momentum conserving $`\delta `$-function, which gives : $`d\sigma {\displaystyle \frac{d^3\stackrel{}{k}_e^{^{}}}{\left(2\pi \right)^3\mathrm{\hspace{0.17em}2}E_e^{}}}{\displaystyle \frac{d^3\stackrel{}{l}}{\left(2\pi \right)^3\mathrm{\hspace{0.17em}2}\mathrm{l}}}{\displaystyle \frac{1}{2E_N^{}}}`$ $`(2\pi )\delta \left(E_e+E_NE_e^{}\sqrt{(\stackrel{}{q}+\stackrel{}{p}_N\stackrel{}{l})^2+M_N^2}\mathrm{l}\right)`$ (A58) $`\times |M_{BORN}|^2(e^2)[{\displaystyle \frac{k_\mu ^{^{}}}{k^{^{}}.l}}{\displaystyle \frac{k_\mu }{k.l}}].[{\displaystyle \frac{k^\mu }{k^{^{}}.l}}{\displaystyle \frac{k^\mu }{k.l}}].`$ Due the energy conserving $`\delta `$-function in Eq. (A58), the integration volume for the soft photon has a complicated ellipsoidal shape in the lab system. In order for the soft-photon phase space integration volume to be spherical, one has to perform the calculation in the c.m. system $`𝒮`$ of the (recoiling nucleon + soft-photon), as discussed in . The system $`𝒮`$ is thus defined by : $`\stackrel{}{p}_N^{^{}}+\stackrel{}{l}=\stackrel{}{q}+\stackrel{}{p}_N=0`$. In the system $`𝒮`$, the energy conserving delta function is independent of the soft-photon angles, and the maximal soft photon energy is isotropic. The integral over the soft-photon momentum (up to some maximum value $`\mathrm{\Delta }E_s`$) can then be performed independently from the integration over the soft photon emission angles. If $`\mathrm{\Delta }E_s`$ is sufficiently small, one can furthermore neglect the soft photon energy with respect to the other energies in the $`\delta `$-function, and perform the integral over the electron momentum $`|\stackrel{}{k}_e^{^{}}|`$ in Eq. (A58). The integration over the outgoing electron momentum eliminates the $`\delta `$-function, which implies the elastic scattering constraint. This yields then for the differential cross section with respect to the outgoing electron angles, the following correction due to soft bremsstrahlung : $$\left(\frac{d\sigma }{d\mathrm{\Omega }_e^{^{}}}\right)_{REALSOFT\gamma }=\left(\frac{d\sigma }{d\mathrm{\Omega }_e^{^{}}}\right)_{BORN}\left(e^2\right)\frac{d^3\stackrel{}{l}}{\left(2\pi \right)^3\mathrm{\hspace{0.17em}2}\mathrm{l}}\left[\frac{k_\mu ^{^{}}}{k^{^{}}.l}\frac{k_\mu }{k.l}\right].\left[\frac{k^\mu }{k^{^{}}.l}\frac{k^\mu }{k.l}\right],$$ (A59) where the soft-photon phase space integral is performed in the system $`𝒮`$, in which the integration volume is spherical. We will denote in the following the external kinematics in the system $`𝒮`$ by tilded quantities ($`\stackrel{~}{E}_e,\stackrel{~}{E}_e^{},\stackrel{~}{E}_N,\stackrel{~}{E}_N^{}`$) to distinguish them from the lab quantities, which we denote by untilded quantities ($`E_e,E_e^{},E_NM_N,E_N^{}`$). To make the transformation between the system $`𝒮`$ and the lab system, we first introduce the missing four-momentum $`p_mp_N^{}+l`$. The system $`𝒮`$ is defined by $`\stackrel{}{p}_m=\stackrel{}{0}`$, and the soft photon limit implies $`p_m^0M_N`$. We can then easily express in the system $`𝒮`$, the energies for the external particles in the elastic scattering process, in terms of lab quantities : $`\stackrel{~}{E}_e`$ $``$ $`{\displaystyle \frac{kp_m}{M_N}}={\displaystyle \frac{1}{M_N}}k(p+q)={\displaystyle \frac{1}{M_N}}(M_NE_eQ^2/2)=E_e^{},`$ (A60) $`\stackrel{~}{E}_e^{}`$ $``$ $`{\displaystyle \frac{k^{}p_m}{M_N}}={\displaystyle \frac{1}{M_N}}k^{}(p+q)={\displaystyle \frac{1}{M_N}}(M_NE_e^{}+Q^2/2)=E_e,`$ (A61) $`\stackrel{~}{E}_N`$ $``$ $`{\displaystyle \frac{pp_m}{M_N}}={\displaystyle \frac{1}{M_N}}p(p+q)=M_N+E_eE_e^{}=E_N^{},`$ (A62) where the elastic scattering condition ($`Q^2=2M_N(E_eE_e^{})`$) has been used in the last step in Eqs. (A60,A61). The angle $`\stackrel{~}{\theta }_e`$ in the frame $`𝒮`$ is obtained from $`k.k^{}=\stackrel{~}{E}_e\stackrel{~}{E}_e^{}(1\mathrm{cos}\stackrel{~}{\theta }_e)=E_eE_e^{}(1\mathrm{cos}\theta _e)`$, which shows (using Eqs. (A60,A61)) that in the soft-photon limit, this angle is the same as in the lab system, i.e. $`\mathrm{cos}\stackrel{~}{\theta }_e=\mathrm{cos}\theta _e`$. The integral of Eq.(A59) extends up to a maximal soft-photon energy $`\mathrm{\Delta }E_s`$ in the system $`𝒮`$, which is expressed in terms of the lab quantities $`E_e`$ and $`E_e^{}`$, by using : $$(p^{}+l)^2M_N^2=(p+kk^{})^2M_N^2=2p(kk^{})+(kk^{})^2,$$ (A63) which leads (for soft-photon energies, i.e. keeping only terms of first order in $`\mathrm{\Delta }E_s`$) to $`2M_N\mathrm{\Delta }E_s`$ $``$ $`2M_N(E_eE_e^{})4E_eE_e^{}\mathrm{sin}^2\theta _e/2,`$ (A64) $`=`$ $`2M_N(E_eE_e^{})2M_N(E_eE_e^{el})E_e^{}/E_e^{el}.`$ (A65) All quantities on the rhs of Eq. (A65) are in the lab, and the elastic scattering condition has been used in the last line ($`E_e^{el}`$ denotes the elastic scattered electron lab energy, to distinguish it from $`E_e^{}`$). From Eq. (A65), one determines then $`\mathrm{\Delta }E_s`$ in terms of lab quantities from the scattered electron spectrum through $$\mathrm{\Delta }E_s=\eta \left(E_e^{el}E_e^{}\right),$$ (A66) where the recoil factor $`\eta `$ is given by $`\eta =E_e/E_e^{el}`$. Deviations from the soft-photon emission formula Eq. (A59) will show up when $`\mathrm{\Delta }E_s`$ is not very small compared with the lepton momenta in the process. The emission of such a semi-hard photon is what is usually referred to as the radiative tail. Although the distinction is somewhat arbitrary, one can always split the integral for photon emission into two parts, one by integrating up to a small value $`\mathrm{\Delta }E_s`$, where the soft-photon approximation in writing down Eq. (A59) holds, and a second integral, starting from this small (but non-zero) value of $`\mathrm{\Delta }E_s`$ up to the energy where one performs the cut in the spectrum. This second integral is finite and can be performed numerically. Such a numerical calculation of the radiative tail without approximations is presented in section IV. In the present section, we give an analytical result for the soft-photon (i.e. small $`\mathrm{\Delta }E_s`$) integral of Eq. (A59), without any further approximations (Remark that in only an approximate evaluation of Eq. (A59) has been given). As is immediately seen by power counting, the integral in Eq. (A59) has a logarithmic IR divergence, corresponding with the emission of photons with zero energy. To demonstrate the cancellation with the IR divergence of the vertex diagram as stated above, one has to regularize the integral of Eq. (A59). In this work this is performed by also using dimensional regularization. The soft photon integral is then evaluated in $`D1`$ dimensions ($`D4`$ corresponds to the physical limit). One now sees that it is extremely advantageous to have a spherical integration volume, in order to evaluate the integral for dimensions $`D4`$. Before continuing the integral of Eq. (A59) into $`D1`$ dimensions, the integration limits for $`l`$ have to be made dimensionless, which leads in the dimensional regularization scheme to the introduction of the same scale $`\mu `$ in Eq. (A67) as was introduced when changing the dimension of the virtual photon loop integral of Eq. (A10). This leads then in $`D1`$ dimensions, to the bremsstrahlung integral : $`I=e^2{\displaystyle ^{\mathrm{l}<\mathrm{\Delta }E_s/\mu }}{\displaystyle \frac{d^{D1}l}{\left(2\pi \right)^{D1}\mathrm{\hspace{0.17em}2}\mathrm{l}}}\left[{\displaystyle \frac{k_\alpha ^{^{}}}{k^{^{}}.l}}{\displaystyle \frac{k_\alpha }{k.l}}\right].\left[{\displaystyle \frac{k^\alpha }{k^{^{}}.l}}{\displaystyle \frac{k^\alpha }{k.l}}\right].`$ (A67) The integral in Eq. (A67) is worked out by introducing polar coordinates in $`D1`$ dimensions. To define the polar angle in the interference term of Eq. (A67), a Feynman parametrization is performed. This leads for $`I`$ to the expression : $`I=`$ $`e^2`$ $`{\displaystyle _0^{\mathrm{l}<\mathrm{\Delta }E_s/\mu }}{\displaystyle \frac{d\mathrm{l}}{\left(2\pi \right)^{D1}}}{\displaystyle \frac{\mathrm{l}^{D2}}{2\mathrm{l}^3}}`$ (A68) $`\times `$ $`{\displaystyle _{D2}}𝑑\mathrm{\Omega }_l\left\{{\displaystyle \frac{k.k^{^{}}}{\stackrel{~}{E}_e\stackrel{~}{E}_e^{^{}}}}{\displaystyle _1^{+1}}𝑑y{\displaystyle \frac{1}{(1\stackrel{~}{\stackrel{}{\beta }_y}.\widehat{l})^2}}{\displaystyle \frac{(1\stackrel{~}{\beta _e}^2)}{(1\stackrel{~}{\stackrel{}{\beta }_e}.\widehat{l})^2}}{\displaystyle \frac{(1\stackrel{~}{\beta _e^{^{}}}^2)}{(1\stackrel{~}{\stackrel{}{\beta }_e^{^{}}}.\widehat{l})^2}}\right\},`$ (A69) where $`\widehat{l}`$ is the unit-vector along the soft photon direction, $`\stackrel{~}{\beta }_e|\stackrel{~}{\stackrel{}{\beta }_e}|`$, $`\stackrel{~}{\beta }_e^{^{}}|\stackrel{~}{\stackrel{}{\beta }_e^{^{}}}|`$ are the incoming and outgoing electron velocities (in the system $`𝒮`$) respectively and where $`\stackrel{~}{\beta }_y|\stackrel{~}{\stackrel{}{\beta }_y}|`$ with $`\stackrel{~}{\stackrel{}{\beta }_e}{\displaystyle \frac{\stackrel{~}{\stackrel{}{k}_e}}{\stackrel{~}{E}_e}},\stackrel{~}{\stackrel{}{\beta }_e^{^{}}}{\displaystyle \frac{\stackrel{~}{\stackrel{}{k}_e^{^{}}}}{\stackrel{~}{E}_e^{^{}}}},`$ (A70) $`\stackrel{~}{\stackrel{}{\beta }_y}\stackrel{~}{\stackrel{}{\beta }_e}{\displaystyle \frac{1}{2}}(1+y)+\stackrel{~}{\stackrel{}{\beta }_e^{^{}}}{\displaystyle \frac{1}{2}}(1y).`$ (A71) The integrals over $`\mathrm{l}`$ and the azimuthal angular integral (over $`D2`$ dimensions) can be performed immediately which yields : $`I=`$ $`e^2`$ $`\left[{\displaystyle \frac{(2\pi )^{2ϵ_{IR}}}{(2\pi )^3}}\left({\displaystyle \frac{\mathrm{\Delta }E_s}{\mu }}\right)^{2ϵ_{IR}}{\displaystyle \frac{1}{4ϵ_{IR}}}\right].\left[{\displaystyle \frac{2\pi }{\pi ^{ϵ_{IR}}}}{\displaystyle \frac{1}{\mathrm{\Gamma }(1ϵ_{IR})}}\right]`$ (A74) $`\times \{{\displaystyle \frac{k.k^{^{}}}{\stackrel{~}{E}_e\stackrel{~}{E}_e^{^{}}}}{\displaystyle _1^{+1}}dy{\displaystyle _1^{+1}}dx{\displaystyle \frac{\left(1x^2\right)^{ϵ_{IR}}}{\left(1\stackrel{~}{\beta _y}x\right)^2}}`$ $`(1\stackrel{~}{\beta _e}^2){\displaystyle _1^{+1}}dx{\displaystyle \frac{\left(1x^2\right)^{ϵ_{IR}}}{\left(1\stackrel{~}{\beta _e}x\right)^2}}(1\stackrel{~}{\beta _e}^{{}_{}{}^{}\mathrm{\hspace{0.17em}2}}){\displaystyle _1^{+1}}dx{\displaystyle \frac{\left(1x^2\right)^{ϵ_{IR}}}{\left(1\stackrel{~}{\beta _e}^{^{}}x\right)^2}}\},`$ The IR divergent term and the finite term are obtained by developing the polar angular integral in Eq. (A74) as $$_1^{+1}𝑑x\frac{\left(1x^2\right)^{ϵ_{IR}}}{\left(1\beta x\right)^2}=_1^{+1}𝑑x\frac{1}{\left(1\beta x\right)^2}ϵ_{IR}_1^{+1}𝑑x\frac{\mathrm{ln}\left(1x^2\right)}{\left(1\beta x\right)^2}+O\left(ϵ_{IR}^2\right).$$ (A75) Performing the integrations in Eq. (A75) (the second integral in Eq. (A75) is simplified by making the substitution $`xu=\beta /(1\beta x)`$ ) yields $$_1^{+1}𝑑x\frac{\left(1x^2\right)^{ϵ_{IR}}}{\left(1\beta x\right)^2}=\frac{2}{1\beta ^2}ϵ_{IR}\frac{2}{1\beta ^2}\left[\mathrm{ln}4+\frac{1}{\beta }\mathrm{ln}\frac{1\beta }{1+\beta }\right]+O\left(ϵ_{IR}^2\right).$$ (A76) Consequently, the IR divergent term and the finite term of the integral $`I`$ are obtained by using Eq. (A76) in Eq. (A74) and by developing all other factors also to order $`ϵ_{IR}`$ : $`I={\displaystyle \frac{e^2}{4\pi ^2}}`$ $`\{[{\displaystyle \frac{1}{ϵ_{IR}}}+\gamma _E\mathrm{ln}{\displaystyle \frac{4\pi \mu ^2}{m^2}}+\mathrm{ln}{\displaystyle \frac{4(\mathrm{\Delta }E_s)^2}{m^2}}][1{\displaystyle \frac{1}{2}}(1\stackrel{~}{\beta }_e\stackrel{~}{\beta }_e^{^{}}\mathrm{cos}\stackrel{~}{\theta }_e)I_y^{(1)}]`$ (A78) $`+[{\displaystyle \frac{1}{2\stackrel{~}{\beta _e}}}\mathrm{ln}{\displaystyle \frac{1\stackrel{~}{\beta }_e}{1+\stackrel{~}{\beta }_e}}+{\displaystyle \frac{1}{2\stackrel{~}{\beta }_e^{^{}}}}\mathrm{ln}{\displaystyle \frac{1\stackrel{~}{\beta }_e^{^{}}}{1+\stackrel{~}{\beta }_e^{^{}}}}{\displaystyle \frac{1}{2}}(1\stackrel{~}{\beta }_e\stackrel{~}{\beta }_e^{^{}}\mathrm{cos}\stackrel{~}{\theta }_e)I_y^{(2)}]\},`$ where the remaining Feynman parameter integrals $`I_y^{(1)}`$ and $`I_y^{(2)}`$ are given by $`I_y^{(1)}{\displaystyle _1^{+1}}𝑑y{\displaystyle \frac{1}{1\stackrel{~}{\beta }_y^2}},`$ (A79) $`I_y^{(2)}{\displaystyle _1^{+1}}𝑑y{\displaystyle \frac{1}{\stackrel{~}{\beta }_y\left(1\stackrel{~}{\beta }_y^2\right)}}\mathrm{ln}{\displaystyle \frac{1\stackrel{~}{\beta }_y}{1+\stackrel{~}{\beta }_y}},`$ (A80) and where $`\stackrel{~}{\beta }_y`$ is given by Eq. (A71). The integral $`I_y^{(1)}`$ in Eq. (A80) can be performed easily and yields $$I_y^{(1)}=\frac{2\stackrel{~}{E}_e\stackrel{~}{E}_e^{^{}}}{m^2}\frac{v^21}{2v}\mathrm{ln}\left(\frac{v+1}{v1}\right),$$ (A81) with $`v`$ as defined in Eq. (A15). To obtain an analytical formula for the integral $`I_y^{(2)}`$ is much harder but was performed in Ref., which we checked Note that the relevant formula quoted in Ref. contains some typing errors. and which yields the result : $`I_y^{(2)}={\displaystyle \frac{1}{|\stackrel{~}{\stackrel{}{\beta }_e}\stackrel{~}{\stackrel{}{\beta }_e^{^{}}}|\mathrm{tanh}\alpha }}`$ $`\{[2\mathrm{ln}(2)+{\displaystyle \frac{1}{2}}\mathrm{ln}(\mathrm{sinh}^2\alpha \mathrm{sinh}^2\varphi _1)]\mathrm{ln}{\displaystyle \frac{\mathrm{sinh}\alpha +\mathrm{sinh}\varphi _1}{\mathrm{sinh}\alpha \mathrm{sinh}\varphi _1}}`$ (A86) $`\mathrm{ln}(\mathrm{sinh}\alpha +\mathrm{sinh}\varphi _1)\mathrm{ln}{\displaystyle \frac{\mathrm{sinh}\alpha \mathrm{sinh}\varphi _1}{4\mathrm{sinh}^2\alpha }}`$ $`+2\mathrm{ln}\left[e^\alpha {\displaystyle \frac{e^\alpha +e^{\varphi _1}}{e^\alpha +e^{\varphi _1}}}\right]\mathrm{ln}{\displaystyle \frac{\mathrm{cosh}\alpha +\mathrm{cosh}\varphi _1}{\mathrm{cosh}\alpha \mathrm{cosh}\varphi _1}}`$ $`\mathrm{\hspace{0.33em}2}\mathrm{\Phi }\left[{\displaystyle \frac{\mathrm{sinh}\alpha +\mathrm{sinh}\varphi _1}{2\mathrm{sinh}\alpha }}\right]+\mathrm{\Phi }\left[\left({\displaystyle \frac{e^\alpha e^{\varphi _1}}{e^\alpha +e^{\varphi _1}}}\right)^2\right]`$ $`\mathrm{\Phi }\left[\left({\displaystyle \frac{e^{\varphi _1}e^\alpha }{e^{\varphi _1}+e^\alpha }}\right)^2\right][\varphi _1\varphi _2]\},`$ where $`\alpha `$, $`\varphi _1`$ and $`\varphi _2`$ are given by : $`\mathrm{cosh}\alpha ={\displaystyle \frac{|\stackrel{~}{\stackrel{}{\beta }_e}\stackrel{~}{\stackrel{}{\beta }_e^{^{}}}|}{\stackrel{~}{\beta }_e\stackrel{~}{\beta }_e^{^{}}\mathrm{sin}\stackrel{~}{\theta }_e}}(\alpha >0),`$ (A87) $`\mathrm{cosh}\varphi _1=\stackrel{~}{\beta }_e\mathrm{cosh}\alpha ,\mathrm{sinh}\varphi _1={\displaystyle \frac{\stackrel{~}{\beta }_e\stackrel{~}{\beta }_e^{^{}}\mathrm{cos}\stackrel{~}{\theta }_e+\stackrel{~}{\beta }_e^2}{\stackrel{~}{\beta }_e\stackrel{~}{\beta }_e^{^{}}\mathrm{sin}\stackrel{~}{\theta }_e}},`$ (A88) $`\mathrm{cosh}\varphi _2=\stackrel{~}{\beta }_e^{^{}}\mathrm{cosh}\alpha ,\mathrm{sinh}\varphi _2={\displaystyle \frac{\stackrel{~}{\beta }_e\stackrel{~}{\beta }_e^{^{}}\mathrm{cos}\stackrel{~}{\theta }_e\stackrel{~}{\beta }_e^{}_{}{}^{}2}{\stackrel{~}{\beta }_e\stackrel{~}{\beta }_e^{^{}}\mathrm{sin}\stackrel{~}{\theta }_e}}.`$ (A89) The function $`\mathrm{\Phi }`$ in Eq. (A86) is given by $$\mathrm{\Phi }(x)\underset{0}{\overset{x}{}}𝑑t\frac{\mathrm{ln}|1t|}{t}.$$ (A90) which agrees with the Spence function (Eq. (A16)) when $`x<1`$. Compared with previous calculations in the literature, it was shown in Ref. that this integral $`I_y^{(2)}`$ was approximated in Ref. and that the calculation of this integral in Ref. contains a factor two error. We also checked the analytical formula of Eq. (A86) by performing the integral of Eq. (A80) numerically. In the ultrarelativistic limit ($`\stackrel{~}{\beta }_e,\stackrel{~}{\beta }_e^{^{}}1`$), the integral $`I_y^{(2)}`$ of Eq. (A86) reduces to $`I_y^{(2)}\stackrel{\stackrel{~}{\beta }_e1,\stackrel{~}{\beta }_e^{^{}}1}{}{\displaystyle \frac{1}{2\mathrm{sin}^2\frac{\stackrel{~}{\theta }_e}{2}}}`$ $`\{{\displaystyle \frac{1}{2}}\mathrm{ln}^2(1\stackrel{~}{\beta }_e^2){\displaystyle \frac{1}{2}}\mathrm{ln}^2(1\stackrel{~}{\beta }_e^{}_{}{}^{}2)+\mathrm{ln}4\mathrm{ln}(1\stackrel{~}{\beta }_e^2)+\mathrm{ln}4\mathrm{ln}(1\stackrel{~}{\beta }_e^{}_{}{}^{}2)`$ (A93) $`+\mathrm{\hspace{0.17em}4}\left(\mathrm{ln}^2\left(\mathrm{sin}{\displaystyle \frac{\stackrel{~}{\theta }_e}{2}}\right)\mathrm{ln}^22\right)\mathrm{\hspace{0.33em}2}\mathrm{ln}\left(\mathrm{cos}^2{\displaystyle \frac{\stackrel{~}{\theta }_e}{2}}\right)\mathrm{ln}\left(\mathrm{sin}^2{\displaystyle \frac{\stackrel{~}{\theta }_e}{2}}\right)`$ $`{\displaystyle \frac{\pi ^2}{3}}\mathrm{\hspace{0.17em}2}Sp\left(\mathrm{sin}^2{\displaystyle \frac{\stackrel{~}{\theta }_e}{2}}\right)\}.`$ Putting all pieces together, the result for the bremsstrahlung cross section accompanying elastic electron scattering is obtained as $`\left({\displaystyle \frac{d\sigma }{d\mathrm{\Omega }_e^{^{}}}}\right)_{REALSOFT\gamma }`$ (A94) $`=\left({\displaystyle \frac{d\sigma }{d\mathrm{\Omega }_e^{^{}}}}\right)_{BORN}\left\{{\displaystyle \frac{\alpha _{em}}{\pi }}\left[{\displaystyle \frac{1}{\epsilon _{IR}}}+\gamma _E\mathrm{ln}\left({\displaystyle \frac{4\pi \mu ^2}{m^2}}\right)\right]\left[{\displaystyle \frac{v^2+1}{2v}}\mathrm{ln}\left({\displaystyle \frac{v+1}{v1}}\right)1\right]+\delta _R\right\},`$ (A95) where the finite part $`\delta _R`$ of the real radiative corrections is given by $`\delta _R`$ $`=`$ $`{\displaystyle \frac{\alpha _{em}}{\pi }}\{\mathrm{ln}\left({\displaystyle \frac{4(\mathrm{\Delta }E_s)^2}{m^2}}\right)[{\displaystyle \frac{v^2+1}{2v}}\mathrm{ln}\left({\displaystyle \frac{v+1}{v1}}\right)\mathrm{\hspace{0.17em}1}]`$ (A97) $`{\displaystyle \frac{1}{2\stackrel{~}{\beta }_e}}\mathrm{ln}\left({\displaystyle \frac{1\stackrel{~}{\beta }_e}{1+\stackrel{~}{\beta }_e}}\right){\displaystyle \frac{1}{2\stackrel{~}{\beta }_e^{^{}}}}\mathrm{ln}\left({\displaystyle \frac{1\stackrel{~}{\beta }_e^{^{}}}{1+\stackrel{~}{\beta }_e^{^{}}}}\right)+{\displaystyle \frac{1}{2}}(1\stackrel{~}{\beta }_e\stackrel{~}{\beta }_e^{^{}}\mathrm{cos}\stackrel{~}{\theta }_e)I_y^{(2)}\},`$ $`\stackrel{Q^2>>m^2}{}`$ $`{\displaystyle \frac{\alpha _{em}}{\pi }}\{\mathrm{ln}\left({\displaystyle \frac{4(\mathrm{\Delta }E_s)^2}{m^2}}\right)[\mathrm{ln}\left({\displaystyle \frac{Q^2}{m^2}}\right)\mathrm{\hspace{0.17em}1}]{\displaystyle \frac{1}{2}}\mathrm{ln}\left({\displaystyle \frac{1\stackrel{~}{\beta }_e^2}{4}}\right){\displaystyle \frac{1}{2}}\mathrm{ln}\left({\displaystyle \frac{1\stackrel{~}{\beta }_e^{}_{}{}^{}2}{4}}\right)`$ (A100) $`{\displaystyle \frac{1}{4}}\mathrm{ln}^2\left(1\stackrel{~}{\beta }_e^2\right){\displaystyle \frac{1}{4}}\mathrm{ln}^2\left(1\stackrel{~}{\beta }_e^{}_{}{}^{}2\right)+\mathrm{ln}2\mathrm{ln}\left(1\stackrel{~}{\beta }_e^2\right)+\mathrm{ln}2\mathrm{ln}\left(1\stackrel{~}{\beta }_e^{}_{}{}^{}2\right)`$ $`+\mathrm{\hspace{0.33em}2}(\mathrm{ln}^2\left(\mathrm{sin}{\displaystyle \frac{\stackrel{~}{\theta }_e}{2}}\right)\mathrm{ln}^22){\displaystyle \frac{\pi ^2}{3}}+Sp\left(\mathrm{cos}^2{\displaystyle \frac{\stackrel{~}{\theta }_e}{2}}\right)\},`$ $`=`$ $`{\displaystyle \frac{\alpha _{em}}{\pi }}\{\mathrm{ln}\left({\displaystyle \frac{(\mathrm{\Delta }E_s)^2}{\stackrel{~}{E}_e\stackrel{~}{E}_e^{^{}}}}\right)[\mathrm{ln}\left({\displaystyle \frac{Q^2}{m^2}}\right)\mathrm{\hspace{0.17em}1}]`$ (A102) $`{\displaystyle \frac{1}{2}}\mathrm{ln}^2\left({\displaystyle \frac{\stackrel{~}{E}_e}{\stackrel{~}{E}_e^{^{}}}}\right)+{\displaystyle \frac{1}{2}}\mathrm{ln}^2\left({\displaystyle \frac{Q^2}{m^2}}\right){\displaystyle \frac{\pi ^2}{3}}+Sp\left(\mathrm{cos}^2{\displaystyle \frac{\stackrel{~}{\theta }_e}{2}}\right)\},`$ where the expression of Eq. (A100) in the $`Q^2>>m^2`$ limit has been rewritten in Eq. (A102) to allow comparison with other expressions found in the literature. Finally to evaluate $`\delta _R`$, we have to express the quantities in the system $`𝒮`$ in terms of lab quantities. The relations given in Eqs. (A60,A61)) yield for elastic scattering : $`\stackrel{~}{E}_e=E_e^{}`$, $`\stackrel{~}{E}_e^{}=E_e`$, and $`\mathrm{cos}\stackrel{~}{\theta }_e=\mathrm{cos}\theta _e`$. From the formula for $`\delta _R`$ (e.g. Eq. (A102), one then sees that one formally obtains exactly the same expression in terms of the lab quantities $`E_e,E_e^{},\theta _e`$. The quantity $`\mathrm{\Delta }E_s`$ is calculated from the cut in the electron spectrum, using the expression of Eq. (A66). A comparison of expressions Eqs. (A95),(A97) with the literature, shows that the same result is obtained as in Ref.. A comparison with the expression used by Mo and Tsai will be given in the next section when we add the vertex correction and soft photon emission contribution, because only their sum is IR finite (and thus independent of the IR regularization procedure used). ### 6 Elastic lepton-nucleon scattering In this section, we bring together the first order radiative corrections at the lepton side (lepton vertex and soft bremsstrahlung from the lepton) and the photon polarization correction to correct the elastic lepton-nucleon scattering cross section. As was shown in the previous sections, these corrections can be calculated model-independently. In the next section, we discuss the additional radiative corrections to the lepton-proton cross section, which originate from the proton side (proton vertex correction, soft bremsstrahlung from proton and two-photon exchange corrections). To calculate these corrections at the proton side, a model for the off-shell (or half off-shell) $`\gamma NN`$ vertex is needed however, and which is therefore to some extent model-dependent. For this latter part, we will refer to the recent work of Ref. . The elastic lepton scattering cross section, corrected to first order in $`\alpha _{em}`$ for the lepton vertex contribution and for the photon polarization contribution, is given by $`\left({\displaystyle \frac{d\sigma }{d\mathrm{\Omega }_e^{^{}}}}\right)_{VIRTUAL\gamma }\left({\displaystyle \frac{d\sigma }{d\mathrm{\Omega }_e^{^{}}}}\right)_{BORN}{\displaystyle \frac{1}{\left(1\stackrel{~}{\mathrm{\Pi }}(Q^2)\right)^2}}\left(1+\mathrm{\hspace{0.33em}2}\left\{F(Q^2)F(Q^2=0)\right\}\right)`$ (A103) $`=`$ $`\left({\displaystyle \frac{d\sigma }{d\mathrm{\Omega }_e^{^{}}}}\right)_{BORN}{\displaystyle \frac{1}{\left(1\stackrel{~}{\mathrm{\Pi }}(Q^2)\right)^2}}`$ (A105) $`\times (1+{\displaystyle \frac{\alpha _{em}}{\pi }}[{\displaystyle \frac{1}{\epsilon _{IR}}}\gamma _E+\mathrm{ln}\left({\displaystyle \frac{4\pi \mu ^2}{m^2}}\right)].[{\displaystyle \frac{v^2+1}{2v}}\mathrm{ln}\left({\displaystyle \frac{v+1}{v1}}\right)\mathrm{\hspace{0.17em}1}]+\delta _{vertex}),`$ where the finite part $`\delta _{vertex}`$ of the lepton vertex correction is found from Eq. (A23) to be given by $`\delta _{vertex}`$ $`={\displaystyle \frac{\alpha _{em}}{\pi }}\{{\displaystyle \frac{v^2+1}{4v}}\mathrm{ln}\left({\displaystyle \frac{v+1}{v1}}\right)\mathrm{ln}\left({\displaystyle \frac{v^21}{4v^2}}\right)+{\displaystyle \frac{2v^2+1}{2v}}\mathrm{ln}\left({\displaystyle \frac{v+1}{v1}}\right)2`$ (A108) $`+{\displaystyle \frac{v^2+1}{2v}}[Sp\left({\displaystyle \frac{v+1}{2v}}\right)Sp\left({\displaystyle \frac{v1}{2v}}\right)]\},`$ $`\stackrel{Q^2>>m^2}{}{\displaystyle \frac{\alpha _{em}}{\pi }}\left\{{\displaystyle \frac{3}{2}}\mathrm{ln}\left({\displaystyle \frac{Q^2}{m^2}}\right)\mathrm{\hspace{0.17em}2}{\displaystyle \frac{1}{2}}\mathrm{ln}^2\left({\displaystyle \frac{Q^2}{m^2}}\right)+{\displaystyle \frac{\pi ^2}{6}}\right\}.`$ In writing down Eq. (A105) to first order in $`\alpha _{em}`$, the contribution of the anomalous magnetic moment term $`G\left(Q^2\right)`$ in the vertex correction Eq. (A9) has been dropped. This contribution vanishes in the ultrarelativistic limit ( $`Q^2>>m^2`$ ) as can be seen from Eq. (A14). The first term in the last line of Eq. (A108) corresponds with the vertex correction term quoted by Mo and Tsai (Eq. ( II.5) of Ref.). The finite part of the photon polarization correction, $`\delta _{vac}2\stackrel{~}{\mathrm{\Pi }}(Q^2)`$, follows from Eq. (A54) as $`\delta _{vac}`$ $`=`$ $`{\displaystyle \frac{\alpha _{em}}{\pi }}{\displaystyle \frac{2}{3}}\left\{\left(v^2{\displaystyle \frac{8}{3}}\right)+v{\displaystyle \frac{\left(3v^2\right)}{2}}\mathrm{ln}\left({\displaystyle \frac{v+1}{v1}}\right)\right\},`$ (A109) $`\stackrel{Q^2>>m^2}{}`$ $`{\displaystyle \frac{\alpha _{em}}{\pi }}{\displaystyle \frac{2}{3}}\left\{{\displaystyle \frac{5}{3}}+\mathrm{ln}\left({\displaystyle \frac{Q^2}{m^2}}\right)\right\},`$ (A110) which agrees with the expression quoted by Mo and Tsai (Eq. ( II.4) of Ref.). To evaluate the vacuum polarization due to $`\mu ^+\mu ^{}`$ and $`\tau ^+\tau ^{}`$ pairs at intermediate $`Q^2`$, one has to use Eq. (A109) instead of the limit of Eq. (A110). Note that an incorrect expression is used in for the vacuum polarization contribution due to $`\mu ^+\mu ^{}`$ pairs (Eq. (A5) in their paper). When adding the real (Eq. (A95)) and virtual (Eq. (A105)) radiative corrections at the lepton side, one verifies that the IR divergent parts exactly cancel. The remaining finite contribution is given to first order in $`\alpha _{em}`$ by $$\left(\frac{d\sigma }{d\mathrm{\Omega }_e^{^{}}}\right)_{VIRTUAL\gamma }+\left(\frac{d\sigma }{d\mathrm{\Omega }_e^{^{}}}\right)_{REALSOFT\gamma }=\left(\frac{d\sigma }{d\mathrm{\Omega }_e^{^{}}}\right)_{BORN}\left(1+\delta _{vac}+\delta _{vertex}+\delta _R\right),$$ (A111) where $`\delta _{vac}`$, $`\delta _{vertex}`$ and $`\delta _R`$ are given by Eqs. (A109), (A108), and (A97)-(A102) respectively. Bringing the three contributions together, leads to the expression (in the $`Q^2>>m^2`$ limit) $`\delta _{vac}+\delta _{vertex}+\delta _R={\displaystyle \frac{\alpha _{em}}{\pi }}`$ $`\{\mathrm{ln}\left({\displaystyle \frac{(\mathrm{\Delta }E_s)^2}{E_eE_e^{^{}}}}\right)[\mathrm{ln}\left({\displaystyle \frac{Q^2}{m^2}}\right)\mathrm{\hspace{0.17em}1}]`$ (A112) $`+`$ $`{\displaystyle \frac{13}{6}}\mathrm{ln}\left({\displaystyle \frac{Q^2}{m^2}}\right){\displaystyle \frac{28}{9}}{\displaystyle \frac{1}{2}}\mathrm{ln}^2\left({\displaystyle \frac{E_e}{E_e^{^{}}}}\right){\displaystyle \frac{\pi ^2}{6}}+Sp\left(\mathrm{cos}^2{\displaystyle \frac{\theta _e}{2}}\right)\},`$ (A113) where $`\mathrm{\Delta }E_s`$, which is the maximum soft-photon energy in the c.m. system of (recoiling proton + soft-photon), is determined as in Eq. (A66), when applying this formula to the scattered electron spectrum. We can compare Eq. (A113) with the recent calculation of Maximon and Tjon , where this calculation was also performed (using a finite photon mass to regularize the IR divergences) without doing any approximations. Comparing Eq. (A113) with their $`Z`$-independent term ($`Z`$ being the hadron charge) - i.e. when not considering radiative corrections at the proton side or two-photon exchange contributions at this point - we find exactly the same result. As was noted in Ref. , the last two terms of Eq. (A113) were omitted by Mo and Tsai . We can approximately take into account the higher order radiative corrections by exponentiating the first order vertex and real radiative corrections. This is strictly true only for the IR divergent part of the vertex correction and soft photon emission contribution, and was demonstrated in Refs. (see e.g. Refs. for pedagogical derivations). The application of this exponentiation procedure also to the finite part consists of an approximation which can be checked by comparing the result with the first order formula of Eq. (A111). For the photon polarization contribution, we iterate the first order vacuum polarization contribution of Eq. (A109) to all orders (resumming all vacuum bubbles of the type of Fig. 19 (c)) by keeping the photon self-energy in the denominator as in Eq. (A105). Remark that a resummation of the first order vacuum polarization contribution does not lead to an exponentiated form. Assuming exponentiation for the finite parts of the vertex and soft photon emission contributions - as occurs for their IR divergent pieces - leads then to the radiative correction formula $$\left(\frac{d\sigma }{d\mathrm{\Omega }_e^{^{}}}\right)_{VIRTUAL\gamma }+\left(\frac{d\sigma }{d\mathrm{\Omega }_e^{^{}}}\right)_{REALSOFT\gamma }=\left(\frac{d\sigma }{d\mathrm{\Omega }_e^{^{}}}\right)_{BORN}\frac{e^{\delta _{vertex}+\delta _R}}{\left(1\delta _{vac}/2\right)^2}.$$ (A114) ### 7 Radiative corrections at the proton side and two-photon exchange contributions In the previous sections, we considered radiative corrections to elastic electron scattering originating solely from the electron side (vertex correction and bremsstrahlung) and from the vacuum polarization. These corrections, which are the dominant ones, can be calculated model independently and follow from QED. To calculate the first order radiative corrections originating from the proton side (proton vertex correction, bremsstrahlung from proton and direct and crossed two-photon exchange contributions), one needs a model for the internal structure of the nucleon because one requires knowledge of off-shell (or half off-shell) $`\gamma NN`$ vertices. This model dependence will become important if one aims at a precision of electron scattering experiments at the 1 % level. To quantify the magnitude of those effects, we refer to the recent work of Maximon and Tjon , where an initial study was performed of the size of internal structure effects. In Ref. , the proton current was taken to have the usual on-shell form and form factors were included in the calculation. The calculation of Ref. goes beyond previous works , as the proton vertex correction and the bremsstrahlung from the proton where calculated without approximations within the given model for the proton current. In the calculation of the direct and crossed box diagrams (two-photon exchange contributions), a less drastic approximation was made in as in (where those box diagrams where only calculated in the soft-photon approximation). The calculation of Ref. yields then the correction formula for elastic electron scattering : $$\left(\frac{d\sigma }{d\mathrm{\Omega }_e^{^{}}}\right)_{TOTAL}=\left(\frac{d\sigma }{d\mathrm{\Omega }_e^{^{}}}\right)_{BORN}(1+\delta _{vac}+\delta _{vertex}+\delta _R+Z\delta _1+Z^2(\delta _2^{(0)}+\delta _2^{(1)}),$$ (A115) where $`\delta _{vac}`$, $`\delta _{vertex}`$ and $`\delta _R`$ are given as above (Eq.(A113)). The terms in Eq. (A115) proportional to $`Z`$ (hadron charge) and $`Z^2`$ contain the corrections from the proton side. The correction $`\delta _1`$, proportional to $`Z`$, was calculated in Ref. as $$\delta _1=\frac{2\alpha _{em}}{\pi }\left\{\mathrm{ln}\left(\frac{4(\mathrm{\Delta }E_s)^2}{Q^2x}\right)\mathrm{ln}\eta +Sp\left(1\frac{\eta }{x}\right)Sp\left(1\frac{1}{\eta x}\right)\right\},$$ (A116) where $`\mathrm{\Delta }E_s`$ and $`\eta `$ are given as in Eq. (A66) and where the variable $`x`$ is defined by $$x=\frac{(Q+\rho )^2}{4M_N^2},\rho ^2=Q^2+4M_N^2,$$ (A117) The correction proportional to $`Z^2`$ was split into two parts in Ref. . The contribution $`\delta _2^{(0)}`$, independent of the nucleon form factors was calculated in Ref. as : $`\delta _2^{(0)}`$ $`=`$ $`{\displaystyle \frac{\alpha _{em}}{\pi }}\{\mathrm{ln}\left({\displaystyle \frac{4(\mathrm{\Delta }E_s)^2}{M_N^2}}\right)({\displaystyle \frac{E_N^{}}{|\stackrel{}{p}_N^{^{}}|}}\mathrm{ln}x1)+\mathrm{\hspace{0.17em}1}`$ (A118) $`+`$ $`{\displaystyle \frac{E_N^{}}{|\stackrel{}{p}_N^{^{}}|}}({\displaystyle \frac{1}{2}}\mathrm{ln}^2x\mathrm{ln}x\mathrm{ln}\left({\displaystyle \frac{\rho ^2}{M_N^2}}\right)+\mathrm{ln}xSp(1{\displaystyle \frac{1}{x^2}})+2Sp({\displaystyle \frac{1}{x}})+{\displaystyle \frac{\pi ^2}{6}})\},`$ (A119) where $`\rho `$ is defined as in Eq. (A117), and where $`E_N^{}`$ ($`|\stackrel{}{p}_N^{^{}}|`$) are the lab energy (momentum) of the recoiling nucleon. For the lengthier expression of $`\delta _2^{(1)}`$, which depends on the nucleon form factors, we refer to Ref. . ## B Treatment of singularities In the numerical calculation of the amplitudes for the virtual photon radiative corrections to the $`epep\gamma `$ reaction, we need to calculate two or three dimensional Feynman parameter integrals, as discussed in section III C. In the integration over the first variable, the numerator consists of polynomials and the denominators may have some structures of the form $`(\alpha ^{}x+\beta ^{}\pm i\epsilon ^{})^n`$, or $`(\alpha ^{}x^2+\beta ^{}x+\gamma ^{}\pm i\epsilon ^{})^n`$ with $`n`$=1,2. Therefore, in the calculations, the following integrals appear : $$\underset{\epsilon ^{}0^+}{lim}_a^b\frac{x^mdx}{(\alpha ^{}x+\beta ^{}\pm i\epsilon ^{})^n}\mathrm{o}r\underset{\epsilon ^{}0^+}{lim}_a^b\frac{x^mdx}{(\alpha ^{}x^2+\beta ^{}x+\gamma ^{}\pm i\epsilon ^{})^n}.$$ (B1) When the denominator has no singularities in the integration range, it is, in principle, easy to calculate these integrals which have the form $$_a^b\frac{x^mdx}{(\alpha ^{}x+\beta ^{})^n}\mathrm{o}r_a^b\frac{x^mdx}{(\alpha ^{}x^2+\beta ^{}x+\gamma ^{})^n}.$$ (B2) Some recurrence relations for these integrals are known , but for small values of $`\alpha ^{}`$ as compared to $`\beta ^{}`$ or to $`\gamma ^{}`$, these relations are unstable numerically. In these cases, we have used either a Taylor expansion or the usual Gauss-Legendre integration method to get very accurate results. In the following part of this appendix, we give the relations used when the denominators in Eq. (B1) have singularities in the integration range except in $`a`$ or $`b`$. The details are given elsewhere . The principle of the method is based on the following relation : $`\underset{\epsilon 0^+}{lim}{\displaystyle _a^b}{\displaystyle \frac{x^mdx}{(xx_0\pm i\epsilon )^n}}`$ (B3) $`=\underset{\epsilon 0^+}{lim}\underset{\eta 0^+}{lim}[{\displaystyle _a^{x_0\eta }}{\displaystyle \frac{x^mdx}{(xx_0\pm i\epsilon )^n}}+{\displaystyle _{x_0\eta }^{x_0+\eta }}{\displaystyle \frac{x^mdx}{(xx_0\pm i\epsilon )^n}}`$ (B4) $`+{\displaystyle _{x_0+\eta }^b}{\displaystyle \frac{x^mdx}{(xx_0\pm i\epsilon )^n}}].`$ (B5) Each integral can be separated in a real part and an imaginary part and we can use for them the analytical expressions given in . Let us start with the case where the denominator is a polynomial of degree 1 in the integration variable. In that case, there is only one singularity for $`x_0=\beta ^{}/\alpha ^{}`$ and the sign of the imaginary part will depend on the sign of $`\alpha ^{}`$. For $`n=1`$ and $`\alpha ^{}>0`$, we have $$\underset{\epsilon ^{}0^+}{lim}_a^b\frac{x^mdx}{\alpha ^{}x+\beta ^{}\pm i\epsilon ^{}}=\frac{1}{\alpha ^{}}\underset{\epsilon 0^+}{lim}_a^b\frac{x^mdx}{xx_0\pm i\epsilon },\epsilon =\frac{\epsilon ^{}}{\alpha ^{}}.$$ (B6) When $`\alpha ^{}<0`$, we have only to replace $`\pm i\epsilon `$ by $`i\epsilon `$ in the right hand side of the Eq. (B6). We now define the following quantities $`J_1={\displaystyle \frac{1}{2}}\mathrm{log}{\displaystyle \frac{(bx_0)^2}{(ax_0)^2}},`$ (B7) $`J_n={\displaystyle \frac{1}{n1}}\left[(bx_0)^{n1}(ax_0)^{n1}\right],n2`$ (B8) to obtain the relations $`\underset{\epsilon 0^+}{lim}{\displaystyle _a^b}{\displaystyle \frac{dx}{xx_0\pm i\epsilon }}=J_1i\pi ,`$ (B14) $`\underset{\epsilon 0^+}{lim}{\displaystyle _a^b}{\displaystyle \frac{xdx}{xx_0\pm i\epsilon }}=J_2+x_0J_1i\pi x_0,`$ $`\underset{\epsilon 0^+}{lim}{\displaystyle _a^b}{\displaystyle \frac{x^2dx}{xx_0\pm i\epsilon }}=J_3+2x_0J_2+x_0^2J_1i\pi x_0^2,`$ $`\underset{\epsilon 0^+}{lim}{\displaystyle _a^b}{\displaystyle \frac{x^3dx}{xx_0\pm i\epsilon }}=J_4+3x_0J_3+3x_0^2J_2+x_0^3J_1i\pi x_0^3,`$ $`\underset{\epsilon 0^+}{lim}{\displaystyle _a^b}{\displaystyle \frac{x^4dx}{xx_0\pm i\epsilon }}=J_5+4x_0J_4+6x_0^2J_3+4x_0^3J_2+x_0^4J_1i\pi x_0^4.`$ $`\mathrm{}`$ For $`n=2`$ and $`\alpha ^{}>0`$, we have $$\underset{\epsilon ^{}0^+}{lim}_a^b\frac{x^mdx}{(\alpha ^{}x+\beta ^{}\pm i\epsilon ^{})^2}=\frac{1}{\alpha _{}^{}{}_{}{}^{2}}\underset{\epsilon 0^+}{lim}_a^b\frac{x^mdx}{(xx_0\pm i\epsilon )^2}.$$ (B15) When $`\alpha ^{}<0`$, we have only to replace $`\pm i\epsilon `$ by $`i\epsilon `$ in the right hand side of the Eq. (B15). We next define the following quantities : $`I_0`$ $`=`$ $`{\displaystyle \frac{1}{ax_0}}{\displaystyle \frac{1}{bx_0}},`$ (B16) $`I_1`$ $`=`$ $`{\displaystyle \frac{1}{2}}\mathrm{log}{\displaystyle \frac{(bx_0)^2}{(ax_0)^2}},`$ (B17) $`I_n`$ $`=`$ $`{\displaystyle \frac{1}{n1}}\left[(bx_0)^{n1}(ax_0)^{n1}\right],n2.`$ (B18) In terms of these quantities, the integrals of Eq. (B1) with $`n=2`$ are given by $`\underset{\epsilon 0^+}{lim}{\displaystyle _a^b}{\displaystyle \frac{dx}{(xx_0\pm i\epsilon )^2}}=I_0,`$ (B19) $`\underset{\epsilon 0^+}{lim}{\displaystyle _a^b}{\displaystyle \frac{xdx}{(xx_0\pm i\epsilon )^2}}=I_1+x_0I_0i\pi ,`$ (B20) $`\underset{\epsilon 0^+}{lim}{\displaystyle _a^b}{\displaystyle \frac{x^2dx}{(xx_0\pm i\epsilon )^2}}=I_2+2x_0I_1+x_0^2I_0i2\pi x_0,`$ (B21) $`\underset{\epsilon 0^+}{lim}{\displaystyle _a^b}{\displaystyle \frac{x^3dx}{(xx_0\pm i\epsilon )^2}}=I_3+3x_0I_2+3x_0^2I_1+x_0^3I_0i3\pi x_0^2,`$ (B22) $`\mathrm{}`$ (B23) We can notice that the real part of these integrals for $`n=1`$ as well as for $`n=2`$ can be derived from the binomial expansion $`(x_0+X)^m`$. In the case of $`n=1`$, the imaginary part is proportional to $`\pi f(x_0)`$ where $`f(x)`$ is the numerator of the integrand. For $`n=2`$, it is straightforward to show that the imaginary part is proportional to $`\pi f^{}(x_0)`$. When the form of the denominator is $`(\alpha ^{}x^2+\beta ^{}x+\gamma ^{}\pm i\epsilon ^{})^n`$, i.e. a polynomial of degree 2 in the integration variable, it is always possible to come back to the preceding cases. When $`\alpha ^{}>0`$, we have $$\underset{\epsilon ^{}0^+}{lim}_a^b\frac{x^mdx}{(\alpha ^{}x^2+\beta ^{}x+\gamma ^{}\pm i\epsilon ^{})^n}=\frac{1}{\alpha _{}^{}{}_{}{}^{n}}\underset{\epsilon 0^+}{lim}_a^b\frac{x^mdx}{(x^2+\beta x+\gamma \pm i\epsilon )^n},$$ (B24) with the following definitions : $$\beta =\frac{\beta ^{}}{\alpha ^{}},\gamma =\frac{\gamma ^{}}{\alpha ^{}},\epsilon =\frac{\epsilon ^{}}{\alpha ^{}}.$$ (B25) The integrand in Eq. (B24) has some singularities when $`\delta =\beta ^24\gamma `$ is positive. When $`\alpha ^{}<0`$, we have only to replace $`\pm i\epsilon `$ by $`i\epsilon `$ on the right hand side of Eq. (B24). It can be shown that $$\underset{\epsilon 0^+}{lim}_a^b\frac{x^mdx}{(x^2+\beta x+\gamma \pm i\epsilon )^n}=\underset{\stackrel{~}{\epsilon }0^+}{lim}_a^b\frac{x^mdx}{(xx_+^R+i\stackrel{~}{\epsilon })^n(xx_{}^Ri\stackrel{~}{\epsilon })^n},$$ (B26) with the definitions $$x_+^R=\frac{\beta +\sqrt{\delta }}{2},x_{}^R=\frac{\beta \sqrt{\delta }}{2},\stackrel{~}{\epsilon }=\frac{2\epsilon }{\delta }.$$ (B27) These integrals can be easily calculated using the decomposition of the fraction into elementary fractions. For $`n=1`$, we obtain $`\underset{\epsilon 0^+}{lim}{\displaystyle _a^b}{\displaystyle \frac{x^mdx}{x^2+\beta x+\gamma \pm i\epsilon }}`$ (B28) $`={\displaystyle \frac{1}{\sqrt{\delta }}}\underset{\stackrel{~}{\epsilon }0^+}{lim}{\displaystyle _a^b}{\displaystyle \frac{x^mdx}{xx_+^R\pm i\stackrel{~}{\epsilon }}}{\displaystyle \frac{1}{\sqrt{\delta }}}\underset{\stackrel{~}{\epsilon }0^+}{lim}{\displaystyle _a^b}{\displaystyle \frac{x^mdx}{xx_{}^Ri\stackrel{~}{\epsilon }}},`$ (B29) and for $`n=2`$ $`\underset{\epsilon 0^+}{lim}{\displaystyle _a^b}{\displaystyle \frac{x^mdx}{(x^2+\beta x+\gamma \pm i\epsilon )^2}}`$ (B30) $`={\displaystyle \frac{1}{\delta }}\underset{\stackrel{~}{\epsilon }0^+}{lim}{\displaystyle _a^b}{\displaystyle \frac{x^mdx}{(xx_+^R\pm i\stackrel{~}{\epsilon })^2}}{\displaystyle \frac{2}{\delta ^{3/2}}}\underset{\stackrel{~}{\epsilon }0^+}{lim}{\displaystyle _a^b}{\displaystyle \frac{x^mdx}{xx_+^R\pm i\stackrel{~}{\epsilon }}}`$ (B31) $`+{\displaystyle \frac{1}{\delta }}\underset{\stackrel{~}{\epsilon }0^+}{lim}{\displaystyle _a^b}{\displaystyle \frac{x^mdx}{(xx_{}^Ri\stackrel{~}{\epsilon })^2}}+{\displaystyle \frac{2}{\delta ^{3/2}}}\underset{\stackrel{~}{\epsilon }0^+}{lim}{\displaystyle _a^b}{\displaystyle \frac{x^mdx}{xx_{}^Ri\stackrel{~}{\epsilon }}}.`$ (B32)
warning/0001/hep-ph0001261.html
ar5iv
text
# The diquark-quark model ## The diquark-quark model <br> The underlying approximations of the diquark–quark model are, first, the neglection of any 3–particle irreducible interactions between the quarks in the nucleon, and, secondly, the separability of correlations in the two–quark channel. The first assumption allows to derive a relativistic Faddeev equation for the 6–point quark Green’s function, and the second one reduces it to an effective quark–diquark BS equation. The full derivation of the quark–diquark BS equation within the NJL model can be found in refs. . In the following we will sketch the most important steps. As usual in the treatment of the BS equation we will work in Euclidean space. Scalar and axialvector diquarks are introduced via the assumption of a separable 4–point quark function: $`G_{\alpha \gamma ,\beta \delta }^{\text{sep}}(p,q,P):=\chi _{\gamma \alpha }(p)D(P)\overline{\chi }_{\beta \delta }(q)+\chi _{\gamma \alpha }^\mu (p)D^{\mu \nu }(P)\overline{\chi }_{\beta \delta }^\nu (q).`$ $`P`$ is the total momentum of the incoming and the outgoing quark–quark pair, $`p`$ and $`q`$ are the relative momenta between the quarks in the two channels. $`\chi _{\alpha \beta }(p)`$ and $`\chi _{\alpha \beta }^\mu (p)`$ are vertex functions of quarks with a scalar and an axialvector diquark, respectively. They belong to a $`\overline{\mathrm{𝟑}}`$–representation in color space and are either flavor antisymmetric (scalar diquark) or flavor symmetric (axialvector diquark). For their Dirac structure we will retain the dominant contribution only. Thus we introduce one scalar function $`P(p)`$ which depends only on the relative momentum $`p`$ between the quarks, $`\chi _{\alpha \beta }(p)`$ $`=`$ $`g_s(\gamma ^5C)_{\alpha \beta }P(p),`$ (3) $`\chi _{\alpha \beta }^\mu (p)`$ $`=`$ $`g_a(\gamma ^\mu C)_{\alpha \beta }P(p).`$ (4) $`C`$ denotes hereby the charge conjugation matrix, and $`g_a`$ and $`g_s`$ are effective coupling constants obtained by normalization of the diquark states. The Pauli principle requires the relative momentum to be defined $`p=\frac{1}{2}(p_\alpha p_\beta )`$, where $`p_\alpha `$ and $`p_\beta `$ are the quark momenta . $`P(p)`$ parametrizes the extension of the vertex in momentum space. To facilitate the comparison between our work and that of we choose $$P(p)=\mathrm{exp}(4\lambda ^2p^2).$$ (5) The propagators of scalar and axialvector diquark are the ones for a free spin-0 and spin-1 particle, $`D(p)`$ $`=`$ $`{\displaystyle \frac{1}{p^2+m_{sc}^2}},`$ (6) $`D^{\mu \nu }(p)`$ $`=`$ $`{\displaystyle \frac{\delta ^{\mu \nu }+(1\xi )\frac{p^\mu p^\nu }{m_{ax}^2}}{p^2+m_{ax}^2}}.`$ (7) $`\xi `$ is a gauge parameter introduced in , and in the following we will put $`\xi =1`$. The constituent quark propagator is simply the free fermion propagator $$S(p)=\frac{ip/m_q}{p^2+m_q^2}.$$ (8) The nucleon BS wave function can be described by an effective multi–spinor characterizing the scalar and axialvector correlations (see e.g. ), $$\mathrm{\Psi }(p,P)u(P)=\left(\begin{array}{c}\mathrm{\Psi }^5(p,P)\\ \mathrm{\Psi }^\mu (p,P)\end{array}\right)u(P)$$ (9) where $`u(P)`$ is a positive–energy Dirac spinor with $`P`$ being the total momentum of the bound state. $`p`$ is the relative momentum between quark and diquark, respectively. The vertex function is defined by amputating the legs off the wave function, $$\left(\begin{array}{c}\mathrm{\Phi }^5\\ \mathrm{\Phi }^\mu \end{array}\right)=S^1\left(\begin{array}{cc}D^1& 0\\ 0& (D^{\mu \nu })^1\end{array}\right)\left(\begin{array}{c}\mathrm{\Psi }^5\\ \mathrm{\Psi }^\nu \end{array}\right).$$ (10) The BS equation is a system of equations for wave and vertex function that takes the compact form $$\frac{d^4p^{}}{(2\pi )^4}G(p,p^{},P)\left(\begin{array}{c}\mathrm{\Psi }^5\\ \mathrm{\Psi }^\mu ^{}\end{array}\right)(p^{},P)=0$$ (11) where the object $`G(p,p^{},P)`$ involves the propagators of quark and diquark and the interaction kernel that describes the quark exchange between quark and diquark, $`G(p,p^{},P)`$ $`=`$ $`(2\pi )^4\delta (pp^{})S^1(p_q)\left(\begin{array}{cc}D^1& 0\\ 0& (D^{\mu \mu ^{}})^1\end{array}\right)(p_d)+`$ (12) $`{\displaystyle \frac{1}{2}}\left(\begin{array}{cc}\chi S^T(q)\overline{\chi }& \sqrt{3}\chi ^\mu ^{}S^T(q)\overline{\chi }\\ \sqrt{3}\chi S^T(q)\overline{\chi }^\mu & \chi ^\mu ^{}S^T(q)\overline{\chi }^\mu \end{array}\right).`$ The partitioning of the momentum between quark and diquark introduces a parameter $`\eta `$ with $`p_q=\eta P+p`$ and $`p_d=(1\eta )Pp`$. Therefore the momentum of the exchanged quark is $`q=pp^{}+(12\eta )P`$. The relative momentum of the quarks at the diquark vertex $`\chi `$ is $`p_2=p+p^{}/2(13\eta )P/2`$ and the one at the conjugated vertex $`\overline{\chi }`$ is $`p_1=p/2+p^{}(13\eta )P/2`$. Relativistic translation invariance requires that if $`\mathrm{\Phi }(p,P;\eta _1)`$ is a solution of the BS equation then also $`\mathrm{\Phi }(p+(\eta _2\eta _1)P,P;\eta _2)`$ is one. The BS equation (11) is solved in the rest frame of the bound state, $`P=\left(\begin{array}{c}\mathrm{𝟎}\\ iM\end{array}\right)`$. In this frame the Salpeter approximation amounts to neglecting the fourth component of all vectors appearing in the interaction, i.e of the vectors $`q`$, $`p_1`$ and $`p_2`$. The immediate consequence is that the vertex function will depend only on the relative three-momentum, $`\mathrm{\Phi }(p,P)\mathrm{\Phi }(𝒑)`$. ## Numerical Solutions <br> In the following we use the complete partial wave decomposition of wave and vertex function for octet baryons given in , $`\left(\begin{array}{c}\mathrm{\Phi }^5(p,P)\\ \mathrm{\Phi }^4(p,P)\\ 𝚽(p,P)\end{array}\right)=\left(\begin{array}{c}\left(\begin{array}{cc}\mathrm{𝟏}S_1& 0\\ \frac{1}{p}(𝝈𝒑)S_2& 0\end{array}\right)\\ \left(\begin{array}{cc}\frac{1}{p}(𝝈𝒑)A_1& 0\\ \mathrm{𝟏}A_2& 0\end{array}\right)\\ \left(\begin{array}{cc}i\widehat{𝒑}(𝝈\widehat{𝒑})A_3+(𝝈\times \widehat{𝒑})(𝝈\widehat{𝒑})A_5& 0\\ \frac{i}{p}𝒑A_4+\frac{1}{p}(𝝈\times 𝒑)A_6& 0\end{array}\right)\end{array}\right).`$ (25) The unknown scalar functions $`S_i`$ and $`A_i`$ are functions of $`p^2=p^\mu p^\mu `$ and of the angular variable $`z=\widehat{P}\widehat{p}`$ which denotes the cosine of the angle between $`p^\mu `$ and the 4-axis. As explained above, the functional dependence collapses in the Salpeter approximation from the two variables ($`p^2,z`$) to the single variable $`p^2(1z^2)`$. We expand the scalar functions in terms of Chebyshev polynomials of the first kind in the variable $`z`$, $$S_i[A_i](p^2,z)=\underset{n=0}{\overset{\mathrm{}}{}}i^nS_i^n[A_i^n](p^2)T_n(z),$$ (26) and derive a system of coupled integral equations for the Chebyshev momenta $`S_i^n[A_i^n]`$ that we solve iteratively as described in . This expansion is very close to the hyperspherical expansion that has been shown to work extraordinarily well in the massive Wick–Cutkosky model and in quenched QED . This finding has been corroborated by in the diquark–quark model for pointlike diquarks. As stated already we adopt the parameters of refs. . In the case of the scalar–axialvector diquark model this especially includes identical diquark normalizations leading to equal diquark–quark–quark couplings in the scalar and axialvector channel, $`g_s=g_a`$. The BS equation (11) can be written as an eigenvalue problem for $`g_s`$. This coupling is adjusted to yield the physical nucleon mass $`M`$=0.939 GeV for given values of the quark and diquark mass as well as the diquark width $`\lambda `$. In table 1 we have listed the two sets of parameters, scalar diquark only and scalar–axialvector diquark model, with their corresponding eigenvalues obtained by us in the full calculation and the Salpeter approximation. The values in parentheses are the ones from refs. . Please note that due to a different flavor normalization these values had to be multiplied by $`\sqrt{2}`$ to be directly comparable to ours. Although we could reproduce the eigenvalue for the model case with scalar diquark only, this is not the case for Set II. We observe that the calculations of involved only 4 instead of 6 axialvector components of $`\mathrm{\Phi }^\mu `$, namely the projected ones onto zero orbital angular momentum and the corresponding lower components. Still one would expect a higher eigenvalue in the reduced system. The more striking observation is the amplification of the eigenvalue by about 20…25% in the Salpeter approximation although the binding energy is small, being only 6% of the sum of the constituent masses. This is in contrast to results obtained in the massive Wick-Cutkosky model where the Salpeter approximation leads to a reduction of the eigenvalue. This may be attributed to the exchange of a boson instead of a fermion as in the present study. The substantial difference between the two approaches is also reflected in the vertex function solutions themselves. Fig. 1 shows the Chebyshev moments of the dominant scalar function $`S_1`$ for both methods using the parameters of Set I. Only the even momenta are given, since the odd ones are zero in the Salpeter approximation (but are present, of course, in the full calculation). Two things are manifest: The Salpeter amplitudes have a much broader spatial extent than the full amplitudes. Secondly, the expansion in Chebyshev polynomials that relies on an approximate $`O(4)`$ symmetry converges much more rapidly for the full solution but is hardly convincing in the Salpeter approximation. Again, since the scalar functions depend in the Salpeter approximation on just one variable, $`p^2(1z^2)`$, our expansion is a cumbersome way of visualizing the solution but makes clear that the Salpeter approximation strongly violates the $`O(4)`$-symmetry. ## Electromagnetic Form Factors <br> The Sachs form factors $`G_E`$ and $`G_M`$ can be extracted from the solutions of the BS equations using the relations $`G_E={\displaystyle \frac{M}{2P^2}}\text{Tr}J^\mu P^\mu ,G_M={\displaystyle \frac{iM^2}{Q^2}}\text{Tr}J^\mu \gamma _T^\mu ,`$ (27) $`J^\mu ={\displaystyle \frac{d^4p_f}{(2\pi )^4}\frac{d^4p_i}{(2\pi )^4}\overline{\mathrm{\Phi }}^T(P_f,p_f)J^\mu \mathrm{\Phi }(P_i,p_i)}.`$ (28) with the definitions $`P=(P_i+P_f)/2`$ and $`\gamma _T^\mu =\gamma ^\mu \widehat{P}^\mu \widehat{P}/`$. To this end, one has to normalize the wave and the vertex function through the condition $$\frac{d^4p}{(2\pi )^4}\frac{d^4p^{}}{(2\pi )^4}\overline{\mathrm{\Psi }}(p^{},P_n)\left[P^\mu \frac{}{P^\mu }G(p^{},p,P)\right]_{P=P_n}\mathrm{\Psi }(p,P_n)\stackrel{!}{=}M\mathrm{\Lambda }^+$$ (29) which uses the definition of $`G`$ given in eq. (12) and employs the positive–energy projector $`\mathrm{\Lambda }^+`$. This normalization integral is again performed easily in the rest frame of the bound state, additionally the double integral over the interaction kernel drops out in the Salpeter approximation since the kernel is independent of $`P`$ in the rest frame. The current operator $`J^\mu `$ in eq. (28) contains all possible couplings of the photon to the kernel $`G`$ of the BS equation. The two simplest contributions make up the impulse approximation pictorially shown in Fig. 2. Since the impulse approximation does not mix scalar and axialvector amplitudes there are altogether four diagrams to compute: quark and diquark coupling and for each of the two the matrix element between scalar and axialvector amplitudes of the nucleon. Flavor algebra yields the following current matrix elements for proton and neutron, $`J_{proton}^{\mu ,\text{imp}}`$ $`=`$ $`{\displaystyle \frac{2}{3}}J_{q,s}^\mu +{\displaystyle \frac{1}{3}}J_{dq,s}^\mu +J_{dq,a}^\mu ,`$ (30) $`J_{neutron}^{\mu ,\text{imp}}`$ $`=`$ $`{\displaystyle \frac{1}{3}}J_{q,s}^\mu +{\displaystyle \frac{1}{3}}J_{dq,s}^\mu +{\displaystyle \frac{1}{3}}J_{q,a}^\mu {\displaystyle \frac{1}{3}}J_{dq,a}^\mu .`$ (31) The diagrams of the impulse approximation separately conserve the current in the Salpeter approximation. The proof depends on the behaviour of the current under time reversal and parity transformation and can be found in and more explicitly in . Furthermore there is the peculiar identity at zero photon momentum transfer, $$J_{q,s[a]}^\mu (Q^2=0)=J_{dq,s[a]}^\mu (Q^2=0),$$ (32) which guarantees that proton and neutron have their correct charge (i.e. $`G_E^p(Q^2=0)=1`$, $`G_E^n(Q^2`$=0)=0). This is not so in the full calculation, to ensure current conservation and the physical charges one has to take into account all diagrams of the photon coupling to the interaction part of $`G`$. These diagrams are shown in fig. 3, the proof of this assertion and the construction of the ”seagull” diagrams (photon coupling to the diquark-quark vertices) can be found in , see also for a general discussion of the current operator in three–body theory. In fig. 4 the electric form factors of proton and neutron are displayed, using parameter set II. The first observation is that in the Salpeter approximation we could not obtain convergence with the expansion in Chebyshev polynomials beyond $`Q^20.4`$ GeV<sup>2</sup>. We computed the form factors in the Breit frame where $`Q`$ is real but $`z_i=\widehat{p}_i\widehat{P}_i`$ and $`z_f=\widehat{p}_f\widehat{P}_f`$ have imaginary parts and their absolute values may exceed one, except for the case of no momentum transfer . So this expansion that works in the rest frame, and it does barely so for the Salpeter approximation, will not generally work in a moving frame. On the other hand, the decrease of the higher Chebyshev moments is so drastic for the full four–dimensional solution that form factor calculations converge easily up to several GeV<sup>2</sup> . However, as can be seen from fig. 4 the Salpeter approximation badly fails above 0.5 GeV<sup>2</sup> thereby revealing its semi–relativistic nature. The second finding concerns the electromagnetic radii. The Salpeter approximation tends to underestimate the proton charge radius and to overestimate the absolute value of the neutron charge radius, see the first two rows of table 3. The axialvectors tend to suppress the neutron electric form factor much more in the full calculation than in the Salpeter approximation. Turning to the magnetic moments, the contributions of the various diagrams are tabulated in table 2 for the proton. Following we ascribed to the axialvector diquark a rather large anomalous magnetic moment of $`\kappa `$=1.6 which was needed by the cited author to fit the proton magnetic moment. As already observed earlier, we could reproduce the magnetic moment for Set I, however, for Set II, the values differ and especially the coupling to the axialvector diquark is much weaker in our Salpeter calculation. Rather more interesting is the comparison in the full calculation between Set I and Set II: The axialvector diquark improves the magnetic moment only marginally! The Salpeter approximation tends to overestimate this contribution quite drastically. Finally we want to mention that the Salpeter approximation underestimates the pion–nucleon coupling $`g_{\pi NN}`$ and the axial coupling $`g_A`$ quite sizeably, see table 3. ## Conclusions <br> In this letter we have presented results for a covariant diquark–quark model. The ladder BS equation for the nucleon has been solved in a fully covariant way and in the instantaneous Salpeter approximation. As for the model with scalar diquarks only we have verified the results of ref. whereas there are discrepancies if the axialvector diquark is included. Part of these differences are due to the fact that in ref. not all (ground state) axialvector components have been taken into account. Additionally, we take our result as an indication that the calculations presented in ref. might suffer from some minor error. The main purpose of this letter is the comparison of observables calculated in the Salpeter approximation to the ones obtained in the fully four–dimensional scheme. The first very surprising observation is the overestimation of the BS eigenvalue in the Salpeter approximation. Phrased otherwise, for a given coupling constant the binding energy would be much too small in the Salpeter approximation. We have also demonstrated that the Salpeter approximation violates badly the approximate $`O`$(4) symmetry of the BS equation. This has drastic consequences for the resulting nucleon electromagnetic form factors if the photon virtuality exceeds 0.4 GeV<sup>2</sup>. Whereas different nucleon radii differ only mildly in these two approaches one sees very clearly that the results for the magnetic moments, the pion–nucleon coupling and the axial coupling are underestimated in the Salpeter approximation. ## Acknowledgement <br> We thank S. Ahlig, H. Reinhardt and H. Weigel for a critical reading of the manuscript and their comments.
warning/0001/cond-mat0001003.html
ar5iv
text
# Luttinger Liquid Physics in the Superconductor Vortex Core ## I Introduction The electronic properties of Type $`II`$ superconductors in the mixed phase are known to have several interesting features. In some pioneering work, Caroli et. al. predicted theoretically the existence of low energy quasiparticle states bound to the core of a single isolated vortex in an $`s`$-wave superconductor. These low lying states are free to move parallel to the vortex, and so give rise to a series of one dimensional bands (*minibands*) as shown in *Fig 1(a)*. The system is still gapped, though the energy gap to the lowest excitation (the minigap) is typically much smaller than the bulk gap. Subsequent theoretical work helped to put this prediction of low energy quasiparticle states bound to the vortex core on firm ground. Striking experimental evidence for the existence of these bound states of the vortex was provided by the scanning tunneling microscopy (STM) studies of Hess et al . These experiments were able to image the vortex quasiparticle states as a function of their energy, and agreed well with theory. Most theoretical work on these low energy quasiparticle states in the vortex core has neglected the effect of the Zeeman coupling between the magnetic field and the spin of the quasiparticles. However as discussed in Ref., and as we argue below, the Zeeman coupling plays an important role in determining the low temperature properties of the system. The Zeeman coupling splits the quasiparticle energy levels, and the minigap is decreased since the energy of one spin species is lowered towards the Fermi energy. For sufficiently strong Zeeman splitting, the minibands of one spin species will be brought down below the Fermi energy, and a filled Fermi sea of spin polarised quasiparticles is formed. We show that the magnetic fields needed in typical materials to form this degenerate quasiparticle system are not large, and could be much smaller than $`H_{c2}`$. We consider the effect of quasiparticle-quasiparticle interactions (ignored in the BCS mean field theory) on the low energy properties of the quasiparticles in the core of the vortex. The quasiparticles are bound in the direction perpendicular to the vortex line but are free to move along it, thus providing an interesting realization of a one dimensional system inside the superconductor. It is well-known in the theory of normal metals that interaction effects are dramatic in one dimension: the generic ground state of the interacting electron system is not a Fermi liquid but a different beast, the Luttinger liquid. We therefore focus attention primarily in the regime of magnetic fields well below $`H_{c2}`$ where the vortices may be treated in isolation. Are some of the striking properties of interacting 1D Fermi systems, such as Luttinger liquid physics, also present in the vortex quasiparticle system? The answer is no, if the interactions and the Zeeman coupling are weak. The presence of the (mini)gap implies that the T=0 state of this system is quite insensitive to weak interactions. However if the Zeeman energy is large enough to start filling the miniband, the system is gapless - interaction effects are then crucial, and we argue that the system is correctly described as a spin polarised Luttinger liquid. We then find that the interaction strength *g* which controls the Luttinger liquid exponents is a function of the miniband filling, and in principle could even be negative. Luttinger liquids have several interesting properties \- the fermion correlation functions are power law with anomalous exponents and the single particle density of states at the Fermi points vanishes as a power law. There have been a small number of experimental realizations of Luttinger liquid behaviour in systems such as 1D semiconductor wires , fractional quantum hall effect edge states and carbon nanotubes . We show that a Luttinger liquid can be realised under appropriate conditions in the vortex core. Tunneling into a Luttinger liquid leads to a characteristic power law tunneling conductance, and we propose an experiment involving Scanning Tunneling Microscopy (STM) measurements on the vortex core as a probe of Luttinger liquid behaviour of the vortex quasiparticles. The quasiparticles also interact with the collective modes of the vortex, which can have interesting consequences. In particular, if inter-vortex interactions (which become increasingly important as we increase the field towards $`H_{c2}`$) are sufficiently strong, we find that a vortex analog of the Peierls effect could occur. The vortex line spontaneously undergoes a periodic modulation of its profile to lower the energy by opening up a gap in the quasiparticle spectrum. The rest of this paper is organised as follows. In Section II, we discuss in a little more detail the core states of an isolated vortex, and the formation of the degenerate quasiparticle gas in the presence of the Zeeman splitting. In Section III we show that interactions between the quasiparticles can lead to them forming a Luttinger liquid, with a varying Luttinger liquid exponent and consider how this state may be observed in STM experiments. In Section IV, we couple the quasiparticles to the vortex collective modes and discuss in particular the possibility of a vortex Peierls transition when inter-vortex interactions are present. Section V discusses the validity of some of our approximations and highlights directions where more work needs to be done. ## II Vortex Quasiparticle States in the Non-Interacting Limit In this section we review the spectrum of the vortex core states and discuss the effect of the Zeeman coupling on it. We begin with the BCS model Hamiltonian in the presence of an arbitrary magnetic field $`\stackrel{}{B}=\stackrel{}{}\times \stackrel{}{A}`$ and pair potential $`\mathrm{\Delta }(r)`$ (we will include the Zeeman term later): $`H_{BCS}`$ $`=`$ $`{\displaystyle d^dxc_\sigma ^{}(x)\epsilon [ieA(x)]c_\sigma (x)}+`$ (2) $`c_{}^{}(x)\mathrm{\Delta }(x)c_{}^{}(x)+c_{}(x)\mathrm{\Delta }^{}(x)c_{}(x)`$ where $`\epsilon [p]`$ is the kinetic energy measured from the chemical potential. For simplicity, we will assume a quadratic energy dispersion $`ϵ(p)=\frac{p^2}{2m}E_F`$. It is convenient to rewrite the Hamiltonian using the change of variables $`d_1(x)`$ $`=`$ $`c_{}(x)`$ $`d_2^{}(x)`$ $`=`$ $`c_{}(x)`$ When written in these variables, there are no anomalous terms in the Hamiltonian: $`H_{BCS}`$ $`=`$ $`{\displaystyle d^dx\left(\begin{array}{cc}d_1^{}(x)& d_2^{}(x)\end{array}\right)H\left(\begin{array}{c}d_1(x)\\ d_2(x)\end{array}\right)}`$ (6) $`H`$ $`=`$ $`\left(\begin{array}{cc}\epsilon [ieA(x)]& \mathrm{\Delta }^{}(x)\\ \mathrm{\Delta }(x)& \epsilon [ieA(x)]\end{array}\right)`$ (9) Physically, the number of $`d`$-particles corresponds to the $`z`$-component of the electron spin. The absence of anomalous terms in the $`d`$ representation thus reflects conservation of the $`z`$-component of the electron spin in the BCS Hamiltonian of Eqn. 2. The Hamiltonian is diagonalised by going over to new quasiparticle operators that are defined as: $`\gamma _{1\alpha }^{}`$ $`=`$ $`{\displaystyle (u_\alpha (x)d_1^{}(x)+v_\alpha (x)d_2^{}(x))d^dx}`$ $`\gamma _{2\alpha }^{}`$ $`=`$ $`{\displaystyle (v_\alpha ^{}(x)d_1^{}(x)u_\alpha ^{}(x)d_2^{}(x))d^dx}`$ The functions $`u_\alpha (x)`$ and $`v_\alpha (x)`$ are found by solving the following eigenvalue equation (with $`E_\alpha 0`$): $$H\left(\begin{array}{c}u_\alpha (x)\\ v_\alpha (x)\end{array}\right)=E_\alpha \left(\begin{array}{c}u_\alpha (x)\\ v_\alpha (x)\end{array}\right)$$ (10) This is just the Bogoliubov deGennes (BdG) equation for the quasiparticle states. The BdG Hamiltonian can be recast in a compact matrix notation as: $$H=\epsilon [i\tau ^𝐳eA(x)]\tau ^𝐳+\mathrm{\Delta }^{}(x)\tau ^++\mathrm{\Delta }(x)\tau ^{}$$ where $`\stackrel{}{\tau }`$ are the $`2\times 2`$ Pauli matrices. In terms of the quasiparticle operators $`\gamma `$, the Hamiltonian is simply: $$H_{BCS}=\underset{\alpha }{}\left(\begin{array}{cc}\gamma _{1\alpha }^{}(x)& \gamma _{2\alpha }^{}(x)\end{array}\right)\left(\begin{array}{cc}E_\alpha & 0\\ 0& E_\alpha \end{array}\right)\left(\begin{array}{c}\gamma _{1\alpha }(x)\\ \gamma _{2\alpha }(x)\end{array}\right)$$ (11) The ground state has all of the negative energy states filled so $`\gamma _{2\alpha }^{}|0=0`$; all positive energy states are unoccupied $`\gamma _{1\alpha }|0=0`$. Note that a spin $``$ excitation is created by ($`\gamma _1^{}`$) while a spin $``$ excitation is created by ($`\gamma _2`$) acting on the ground state. After these generalities, let us specialise to the quasiparticle states bound to the vortex line in a superconductor. Consider an isolated, straight vortex oriented along the z-axis, sitting in a clean, conventional Type II superconductor. This is realised if the applied magnetic field is just above $`H_{c1}`$. The vortex bound state energy levels are obtained by solving Eqn. 10 with the appropriate gap profile ($`\mathrm{\Delta }(r)=|\mathrm{\Delta }(r)|e^{i\varphi }`$, where $`\varphi `$ is the azimuthal angle about the vortex). It is useful to perform the singular gauge transformation that makes $`\mathrm{\Delta }`$ real everywhere (the London gauge): $`\left(\begin{array}{c}u^{}\\ v^{}\end{array}\right)`$ $`=`$ $`\left(\begin{array}{c}e^{+i\frac{\varphi }{2}}u\\ e^{i\frac{\varphi }{2}}v\end{array}\right)`$ (16) $`H^{}\left(\begin{array}{c}u^{}\\ v^{}\end{array}\right)`$ $`=`$ $`E\left(\begin{array}{c}u^{}\\ v^{}\end{array}\right)`$ (21) $`H^{}=[{\displaystyle \frac{1}{2m}}(i\tau ^𝐳eA{\displaystyle \frac{\widehat{\varphi }}{2r}})^2`$ $``$ $`E_F]\tau ^𝐳+|\mathrm{\Delta }(r)|\tau ^𝐱`$ (22) The transformed quasiparticle wavefunctions $`\left(\begin{array}{c}u^{}\\ v^{}\end{array}\right)`$ obey antiperiodic boundary conditions to keep $`\left(\begin{array}{c}u\\ v\end{array}\right)`$ single valued. We first review the results of Ref. on the structure of an isolated vortex. In Ref. the effect of the magnetic field on the low energy quasiparticles is not included. This is only a good approximation for an isolated vortex in an extreme Type II superconductor; then the flux enclosed in the spatial region where the low energy states are bound is a small fraction of the flux quantum, and hence the magnetic field has little effect on these states. Subsequently we will see how these results need to be modified on including the effect of the magnetic field. The low energy spectrum of the vortex core states then is : $$E^0(\mu ,k_z)=\mu \left(\frac{c\mathrm{\Delta }_{\mathrm{}}}{k_F\xi }\right)\left[1\frac{k_z^2}{k_F^2}\right]^{\frac{1}{2}}$$ (23) and is shown in Fig 1(a). The corresponding quasiparticle wavefunctions in cylindrical coordinates take the form $`\left(\begin{array}{c}u_{\mu k}^{}(r,\varphi ,z)\\ v_{\mu k}^{}(r,\varphi ,z)\end{array}\right)`$ $`=`$ $`e^{i\mu \varphi }e^{ik_zz}\left(\begin{array}{c}f_{\mu k_z}(r)\\ g_{\mu k_z}(r)\end{array}\right)`$ (28) where $`\mu \{\frac{1}{2},\frac{3}{2},\frac{5}{2}\mathrm{}\}`$ appears as an angular momentum in the wavefunctions and is a half integer due to the anti-periodic boundary conditions, $`k_z`$ is the momentum along the vortex line, $`c`$ is a constant of order one and $`\xi `$ is the zero temperature coherence length. We have denoted by $`\mathrm{\Delta }_{\mathrm{}}`$ the value of the gap very far away from the vortex. The $`k_z`$ dependence of the energy comes from the motion along the vortex line. The energy scale of the minigap is $`\frac{\mathrm{\Delta }_{\mathrm{}}^2}{E_F}`$ and is much smaller than the bulk gap $`\mathrm{\Delta }_{\mathrm{}}`$. The radial wave funcions $`f`$ and $`g`$ are mainly confined to within a radius $`\xi `$ (the coherence length) about the vortex for the low energy states ($`E(\mu ,k_z)\mathrm{\Delta }_{\mathrm{}}`$). Now consider the effect of the magnetic field $`\stackrel{}{B}=B\widehat{z}`$. We assume though that the field is much less than $`H_{c2}`$ so that the vortices are well separated. Two effect occur : (a) the Zeeman term splits the energy levels and (b) there is an orbital effect. We consider these in turn, starting with the orbital effect. For the magnetic fields of interest, the inter-vortex separation is typically smaller than the penetration depth so that the field in the superconductor is fairly uniform and equal to the external field ($`=B\widehat{z}`$, and we take $`\stackrel{}{A}(\stackrel{}{x})=\frac{B}{2}(y,x,0)`$). In the presence of the vector potential additional terms are generated in the BdG Hamiltonian which in the London gauge take the form: $`\delta H`$ $`=`$ $`\delta H_1+\delta H_2`$ (29) $`\delta H_1`$ $`=`$ $`{\displaystyle \frac{|e|}{2m}}(A(x)p+pA(x))`$ (30) $`\delta H_2`$ $`=`$ $`{\displaystyle \frac{(eA(x))^2}{2m}}\tau ^𝐳+{\displaystyle \frac{|e|B}{2m}}\tau ^𝐳`$ (31) $`\delta H_2`$ may be neglected since the first term is small compared to the minigap for fields well below $`H_{c2}`$ and the second term only makes a small shift of the Fermi energy. So we are left with $`\delta H_1`$ which can be written as: $$\delta H_1=\frac{|e|}{2m}B\stackrel{}{r}\times \stackrel{}{p}=\frac{|e|}{2m}BL_z$$ (32) This does not affect the eigenstates (28) but shifts their energy by $$E(\mu ,k_z,B)=E^0(\mu ,k_z)+(\frac{|e|B}{2m})\mu $$ (33) The orbital effect therefore increases the magnitude of energy of the quasiparticle states. The Zeeman term takes the form $`H_Z`$ $`=`$ $`{\displaystyle \frac{g\mu _BB}{2}}{\displaystyle (c_{}^{}(x)c_{}(x)c_{}^{}(x)c_{}(x))d^dx}`$ (34) $`=`$ $`{\displaystyle \frac{g\mu _BB}{2}}{\displaystyle \left(d_1^{}(x)d_1(x)+d_2^{}(x)d_2(x)\right)d^dx}`$ (35) $`=`$ $`{\displaystyle \frac{g\mu _BB}{2}}{\displaystyle \underset{\alpha }{}}(\gamma _{1\alpha }^{}\gamma _{1\alpha }+\gamma _{2\alpha }^{}\gamma _{2\alpha })`$ (36) (In the second line, we have dropped an irrelevant constant and $`\mu _B=\frac{|e|\mathrm{}}{2m}`$ is the Bohr magneton). The Zeeman term thus behaves as a ‘chemical potential’ for the d-particles . Increasing the magnetic field raises this chemical potential. Beyond a certain field it enters the first miniband and a degenerate Fermi sea of quasiparticles forms as shown in Fig 1(b). The condition for this is simply: $$E(\mu =\frac{1}{2},k_z=0,B)\frac{g\mu _BB}{2}$$ (37) or equivalently $$E^0(\mu =\frac{1}{2},k_z=0)\frac{(g1)}{2}\mu _BB$$ (38) Thus a degenerate gas of spin polarised fermions (in this case all spin $``$) can form due to the Zeeman coupling if the magnetic field exceeds $`H_{cZ}=\frac{2}{\mu _B}E^0(\mu =\frac{1}{2},k_z=0)`$ (assuming $`g=2`$). Note that it is only the first miniband that can be brought below the chemical potential. For $`\mu >\frac{1}{2}`$, the orbital effect wins over the Zeeman splitting, and the levels do not cross the chemical potential. The magnetic fields needed to begin filling the miniband states are very reasonable, since the minigap energy is typically small. To illustrate this point we consider the case of NbSe<sub>2</sub> for which numerical calulations of the vortex quasiparticle spectrum are available . NbSe<sub>2</sub> has a superconducting transition temperature of $`T_c=7.2K`$, and numerical calculations find the minigap to be $`0.4`$ Kelvin. The magnetic field required to close the minigap is \[$`H_{cZ}`$\] is $`1.3`$ Tesla (we have assumed $`g=2`$) while $`H_{c2}`$ is larger at $`3.2`$T for this material. Better candidate materials will have smaller ratios of $`\frac{H_{cZ}}{H_{c2}}`$ than NbSe<sub>2</sub>. In passing we note that the Zeeman splitting gives rise to quasiparticles even at zero temperature, and in doing so ‘melts’ some of the condensate at the centre of the vortex. The new profile $`\mathrm{\Delta }(r)`$, can be calculated from the self consistency equation. This is analogous to the Kramer-Pesch effect , where temperature does the job of melting the condensate and expanding the vortex. We shall not take into account the effects of the changing gap profile on the quasiparticle states, as they are expected to be small. ## III Interaction Effects and Luttinger liquid Formation Interactions between fermions in clean 1D systems have a dramatic effect resulting in Luttinger liquid behaviour. Here we shall consider the effect of interactions between the quasiparticles in the vortex core that form the 1D Fermi gas. A few comments on the nature of these interactions is in order. The underlying microscopic electronic system is characterized, at low energies, by three qualitatively different kinds of interactions: (a) the spin singlet density-density interaction in the particle-hole channel (b) the triplet spin density-spin density interaction in the particle-hole channel and (c) the interaction in the particle-particle channel. Conventional BCS superconductors arise when the interactions in the particle-particle channel are attractive. Indeed the electron gas is unstable toward the formation of Cooper pairs in the presence of such an attractive interaction. The BCS theory simply treats this attractive interaction in the particle-particle channel in a mean field approximation. One still however has to deal with the interactions in the particle-hole channel. (These are simply ignored in the so-called reduced BCS Hamiltonian). Inclusion of these leads to interactions between the quasiparticles in the superconductor. Note that the long-ranged Coulomb repulsion between the electrons in the singlet channel is screened out by the condensate in the superconducting phase. Thus the residual interactions between the quasiparticles are short-ranged. A further source of interactions between the quasiparticles comes from the inclusion of fluctuations of the mean field order parameter (again usually ignored in BCS theory). Our general conclusions are insensitive to the precise form of this short-ranged interaction between the quasiparticles. We will therefore, for illustrative purposes, consider a particular model interaction. We add to the BCS Hamiltonian Eqn.2 the “interaction” term $$H_{int}=\underset{\sigma \sigma ^{}}{}V(|xx^{}|)c_\sigma ^{}(x)c_\sigma ^{}^{}(x^{})c_\sigma ^{}(x^{})c_\sigma (x)d^3xd^3x^{}$$ (39) where $`V(|xx^{}|)`$ is assumed to be short-ranged. To study the effect on the vortex core states, it is convenient to re-express this in terms of the $`\gamma `$ operators introduced in Section II $`H_{int}`$ $`=`$ $`{\displaystyle \underset{\alpha \beta \alpha ^{}\beta ^{}}{}}{\displaystyle [V(|xx^{}|)]}:(u_\alpha ^{}(x)\gamma _{1\alpha }^{}+v_\alpha (r)\gamma _{2\alpha }^{})`$ (44) $`\times (u_\beta (x)\gamma _{1\beta }+v_\beta ^{}(x)\gamma _{2\beta })`$ $`\times (v_\alpha ^{}^{}(x^{})\gamma _{1\alpha ^{}}^{}u_\alpha ^{}(x^{})\gamma _{2\alpha ^{}}^{})`$ $`\times (v_\beta ^{}(x^{})\gamma _{1\beta ^{}}u_\beta ^{}^{}(x^{})\gamma _{2\beta ^{}}):d^dxd^dx^{}`$ $`+\mathrm{}+\mathrm{}+\mathrm{}`$ where the normal ordering is with respect to the superconductor vacuum, and we have written out only the first of the four terms present from different spin combinations. Let us study the effect of interactions when we may treat the vortices as isolated and the lowest miniband is occupied with the two Fermi points at $`\pm k_f`$ (so $`E_{\frac{3}{2},0}>E_{Zeeman}>E_{\frac{1}{2},0}`$). Notice that the Fermi Sea consists of spin up quasiparticles only. The excitation spectrum for this case contains a low energy part of ‘particle-hole’ excitations in the vicinity of the Fermi points, and a high energy sector that involves either $`\gamma _1`$ quasiparticles hopping into higher minigap states, or the destruction of $`\gamma _2`$ quasiparticles from deep below the chemical potential. The latter two are gapped, with the gap being of order the minigap energy ($`\frac{\mathrm{\Delta }}{k_F\xi }`$). Since we are interested in the effects of weak interactions on the $`T=0`$ state of the system, we confine ourselves to the low energy sector of the problem. Formally, we can achieve this by defining a projection operator $`P_ϵ`$ that only retains states close to the Fermi points, i.e. $`\{\gamma _1`$ quasiparticles, $`\mu =\frac{1}{2}`$, $`k_z(k_fϵ,k_f+ϵ)`$ or $`k_z(k_fϵ,k_f+ϵ);ϵk_f\}`$. Now, we can project the interaction term Eqn(44) down into this subspace, and it is easy to see that the only non-trivial term that remains is: $$H_{int}^P=U\underset{|q|<ϵ}{}\rho _L(q)\rho _R(q)$$ (45) where the left (and similarly the right) density operators $`\rho _L`$ and $`\rho _R`$ are constructed from the fermion operators in the usual way, $`\rho _L`$ for example is given by: $$\rho _L=\underset{|k+k_f|<ϵ}{}\mathrm{\Gamma }_{k+q}^{}\mathrm{\Gamma }_k$$ and $`\mathrm{\Gamma }_k=\gamma _{1,\frac{1}{2},k}`$. In general, doing a quantitative calculation of $`U`$ for a real material is not an easy task because a complete knowledge of the interaction potential and quasiparticle wavefunctions is needed. We simply note that $`U`$ will be a function of the filling of the miniband ($`U=U(k_f)`$), and in principle can even be negative (attractive interaction between the quasiparticles) . For the particular interaction of Eqn. 39 we can express $`U`$ in terms of the interaction potential and the wavefunction of the states near the Fermi points: $$U=_{x,y,x^{},y^{}}V(xx^{},yy^{},z)Q_{k_f}(x,y)Q_{k_f}(x^{},y^{})[1e^{2ik_fz}]$$ (46) Here the quasiparticle ‘charge’ $`Q_{k_f}`$ is defined by $$Q_{k_f}(x,y)=|u_{\frac{1}{2},k_f}(x,y)|^2|v_{\frac{1}{2},k_f}(x,y)|^2$$ (47) We thus have a system of one-dimensional fermions with two Fermi points interacting through $`H_{int}^P`$. The fermions are spin-polarized (and hence, effectively, spinless). It is well-known that such a system is correctly described not as a Fermi liquid, but rather as a Luttinger liquid. Luttinger liquids are characterised by power law correlations. The exponents are controlled by the dimensionless quantity $`g=\frac{U}{2\pi v_f}`$ where $`v_f`$ is the Fermi velocity of the one dimensional system. For example, $`\mathrm{\Gamma }^{}(z)\mathrm{\Gamma }(z^{})(1/|zz^{}|^\nu )`$ where $`z`$ is the position along the vortex line and the exponent $`\nu =(1/\sqrt{1g^2})`$. Relatedly, tunneling into the bulk of a Luttinger liquid is characterised by nonlinear I-V characteristics, in fact at low voltages the tunneling current-voltage relation is of the form $`IV^\nu `$. This suggests that STM could be used to experimentally verify Luttinger liquid formation in the vortex cores under the conditions described above. At sufficiently low temperatures when the Luttinger liquid behaviour is dominant, we expect the tunneling conductance close to the centre of the vortex to be given by : $$\sigma (V,x)|u_{\frac{1}{2},k_f}(x)|^2|V|^{\nu _e1}$$ (48) and $`\nu _e`$ is the exponent for tunneling into the edge of a Luttinger liquid $`\nu _e=(1+g)^{\frac{1}{2}}(1g)^{\frac{1}{2}}`$ . The temperatures at which such Luttinger liquid behaviour will be observed is necessarily small compared to the degeneracy temperature, which for the half filled miniband of $`NbSe_2`$ is $`0.4`$Kelvin . These effects may be observed at higher temperatures by using a superconductor with a larger minigap, but this would also require larger magnetic fields to close the minigap via the Zeeman splitting. ## IV Coupling to Vortex Collective Modes and Vortex Peierls Effect So far we have only considered straight and rigid vortex lines, however in reality the vortex is a soft object and can undergo shape fluctuations. These collective modes of the vortex interact with the quasiparticles bound to the vortex line. As we show below the interacting quasiparticle - collective mode system is similar to the problem of interacting electrons and phonons in one dimension, which is known to undergo a Peierls transition (for commensurate filling) in which the lattice spontaneously distorts and a gap opens at the Fermi points. The gain in electronic energy by opening of the gap offsets the elastic energy cost of the lattice distortion. The electron scattering from the vicinity of one Fermi point to the other (2$`k_F`$ scattering) involves vanishing energy denominators and is responsible for this instability. We can ask if an analogous effect occurs in the degenerate 1D quasiparticle system in the vortex core interacting with the shape fluctuations of the vortex - which can give rise to 2$`k_F`$ scatterings . Note that for an isolated vortex line, which is a one dimensional object it is not possible to spontaneously break the continuous symmetry of translations along the vortex line. However in the vortex lattice if inter vortex interactions are sufficiently strong, a vortex analog of the Peierls effect could occur at finite temperature. This issue was raised earlier by Bouchaud who modelled the vortex core simply as a wire of normal electrons. However we believe this is not an adequate model to discuss this phenomena. A degenerate Fermi system is formed in the vortex only in the presence of the Zeeman splitting. The collective modes of the vortex may be classified according to their angular momentum. Scalar modes scatter quasiparticles within the same $`\mu `$ miniband, while the lowest order coupling of quasiparticles to collective modes of higher angular momenta (say $`l`$) will scatter quasiparticles from one miniband ($`\mu `$) to a different one ($`\mu \pm l`$) by conservation of angular momentum. Since the low energy excitations all lie in a single $`\mu =\frac{1}{2}`$ miniband we only consider coupling of quasiparticles to the scalar collective modes of the vortex. Scalar vortex modes modulate the radial profile of the vortex as one moves along the vortex line. For instance, a periodic modulation of the vortex radius (defined say as when the order parameter reaches half its asymptotic value) is an example of a scalar deformation of the vortex. If the scalar normal modes of the vortex are labelled by $`n`$, let $`\sigma _n^{}(q_z)`$ create the n<sup>th</sup> radial mode with momentum $`q_z`$ along the vortex line. The coupling of the vortex modes to the quasiparticles (of the partially filled miniband) can be derived from the BdG equation, or simply from symmetry is found to be: $$H_{int}^{scalar}=\underset{q,k,n}{}g_{int}(k,q,n)\sigma _n(q)\mathrm{\Gamma }_{k+q}^{}\mathrm{\Gamma }_k+h.c.$$ (49) The Hamiltonian for the scalar modes is: $$H_{modes}=\underset{n,q}{}E(n,q)\sigma _n^{}(q)\sigma _n(q)$$ (50) Thus, scalar modes of the vortex of momentum 2$`k_f`$ will scatter quasiparticles from one Fermi point to the other. As we have noted previously, for an isolated vortex this does not lead to a gapped phase and the Luttinger liquid behaviour will persist. However if inter vortex interactions are sufficiently strong, then a finite temperature transition to a vortex Peierls phase can occur. Which of the modes $`n_0`$ has the highest transition temperature is a function of $`g_{int}`$, energy and inter vortex coupling for that mode. In the vortex Peierls phase, the quasiparticle spectrum is gapped, and we have a collective mode condensate i.e. $`\sigma _{n_0}(2k_f)0`$ which implies a periodic modulation of the vortex profile. We do not attempt to predict the precise conditions for the experimental realization of this phase. Increasing the vortex density, and hence increasing intervortex interactions can raise the transition temperature, as long as the core quasiparticles do not hop between vortices and destroy the one-dimensional nature of the system. Within a mean field approximation, the Fermi gas always undergoes the Peierls transition. The instability is enhanced for repulsive Luttinger liquids . Thermal transport measurements along the vortex line are expected to be sensitive to the vortex Peierls transition. The Luttinger liquid behaviour of the vortex quasiparticles would also be destroyed if pairing of quasiparticles were to occur. This would induce an additional (eg. p-wave) pairing amplitude in the vortex core, and could also lead to breaking of the rotational symmetry of the vortex line. This interesting situation was considered by Makhlin and Volovik in . Once again, the transition temperature for this instability is nonzero only in the presence of suitable inter-vortex interactions. ## V Discussion We have shown that a one-dimensional degenerate gas of quasiparticles can form in the vortex core, as a result of the Zeeman coupling and can serve as a laboratory for one dimensional physics. We showed that quasiparticle interactions drive the formation of a Luttinger liquid inside the vortex core. The interaction strength $`U`$ of the Luttinger liquid was shown to be a function of the filling, and could even be negative. Thus we have a 1D interacting Fermi system, where novel features arise due to the fact that the fermions involved are superconducting quasiparticles and the ‘wire’ confining them is a vortex. We now examine in more detail some of the approximations that have been made, and the prospects for verifying the results we predict in experiments. One of the main approximations that we have made is to treat each vortex as effectively isolated, which gives rise to the one dimensional nature of the vortex quasiparticle system. As the magnetic field is increased towards $`H_{c2}`$, the vortices get closer to each other and the wavefuctions of the vortex core states start to overlap. This leads to quasiparticle hopping between vortices that will eventually destroy the one dimensional nature of the system. To keep the hopping small, we need that the separation between vortices at the magnetic fields of interest is much larger than the coherence length. We therefore require that the field $`H_{cZ}`$ at which levels start to fill satisfies $`H_{cZ}H_{c2}`$. However, the temperature scale associated with the field $`H_{cZ}`$ should not be too small since the physics of interest occurs only below that temperature, suggesting materials with a relatively high $`T_c`$. A possible family of candidates that meet these criteria are the borocarbides . Another possibility is to work with a material that has $`H_{cZ}<H_{c1}`$. Then even at magnetic fields just above $`H_{c1}`$, when the vortices are very well isolated, a degenerate quasiparticle liquid will be formed in the vortex cores. Elemental Niobium, that has a relatively large $`H_{c1}=0.14`$T is possibly such a system. Throughout this work we have assumed that the superconductor is perfectly clean, but in any real system disorder is always present. The stability of the Luttinger liquid to weak disorder depends on the value of the interaction strength $`g`$ . It is well-known that there exists a critical $`g_c<0`$ such that for $`g>g_c`$, any disorder kills the Luttinger liquid behaviour leading instead to a phase with localized quasiparticle excitations. Still, for sufficiently clean systems there is a range of temperatures for which the properties of the system are controlled by the LL fixed point, even though the ultimate zero temperature state may be localised. In that case, the scanning tunneling conductance that we predict for the LL should be obtained in this crossover regime. On the other hand, if $`g<g_c`$, the Luttinger liquid is, in fact, stable to weak disorder. Since $`g`$ for the vortex core Luttinger liquid could be negative, the following interesting effect can occur. By varying the external magnetic field, $`g`$ can be varied and the system can be tuned through a one dimensional delocalization transition from the phase with localized quasiparticles (at small negative $`g`$) to the Luttinger liquid (at large negative $`g`$). The formation of the degenerate quasiparticle gas in the vortex core can be probed by measurements of the low temperature specific heat - a linear temperature dependence with a field dependent coefficient should obtain at the lowest temperatures. Once signatures of the formation of this quasiparticle gas are observed and some of its physical parameters (eg. the minigap and dispersion with $`k_z`$) measured, it would be of great interest to look for possible Luttinger liquid behaviour in, for instance, STM tunneling into the vortex core. We also considered coupling the vortex quasiparticles to the collective modes of the vortex. We find that a vortex analog of the Peierls transition with a non-zero transition temperature could arise if the inter-vortex interactions are strong enough. ## VI Acknowledgements We would like to thank K. Damle, C. Dasgupta, D. Huse, N.P. Ong, A. Melikidze, S. Sondhi and especially D. Haldane for useful discussions. One of us (A.V.) acknowledges support from grant NSF DMR-9809483. T.S was supported by NSF Grants DMR-97-04005, DMR95-28578 and PHY94-07194.
warning/0001/nucl-th0001052.html
ar5iv
text
# FZJ-IKP(TH)-2000-02 Ordinary and radiative muon capture on the proton and the pseudoscalar form factor of the nucleon ## I Introduction With the ubiquitous success of the Standard Model<sup>#4</sup><sup>#4</sup>#4For a recent review of utilizing various muon capture reactions as filters for physics beyond the Standard Model see . ordinary and radiative muon capture on the proton can nowadays be considered as an excellent testing ground for our understanding of spontaneous and explicit chiral symmetry breaking in QCD. This stems from the fact that the typical momentum transfer in these reactions is very small—of the order of the muon mass—and one therefore can apply effective field theory methods, in particular baryon chiral perturbation theory. Ordinary muon capture (OMC), $$\mu ^{}(l)+p(r)\nu _\mu (l^{})+n(r^{}),$$ (1.1) where we have indicated the four–momenta of the various particles, allows to measure the so–called induced pseudoscalar coupling constant, $`g_P`$. This coupling constant is nothing but the value of the induced pseudoscalar form factor $`G_P(t)`$ at the four–momentum transfer for muon capture by the proton at rest, $`g_p=m_\mu G_P(t=0.88m_\mu ^2)/2M_N`$, with $`m_\mu (M_N)`$ the muon (nucleon) mass. Theoretically, it is dominated by the pion pole as given by the time–honored PCAC prediction, $`g_P^{\mathrm{PCAC}}=8.89`$. Adler and Dothan as well as Wolfenstein calculated the first correction to the PCAC result utilizing by now outdated (and in some cases misleading) methods. The ADW relation was rederived directly from QCD Ward identities within the framework of heavy baryon chiral perturbation theory and shown not to be affected by $`\mathrm{\Delta }(1232)`$ effects in , $`g_P^{\mathrm{CHPT}}=8.44\pm 0.23`$ (a similar calculation with a slightly different result was given in ). The presently available data have, however, too large error bars to discriminate between the pion pole prediction and its corrected version . Furthermore, $`G_p(t)`$ for low four–momentum transfer squared can be extracted from charged pion electroproduction measurements . The resulting momentum dependence of the induced pseudoscalar form factor is in agreement with the pion pole prediction. However, as in the case of the pseudoscalar coupling, the data are not precise enough to be sensitive to the small corrections found in . While the momentum transfer in OMC is fixed, radiative muon capture (RMC), $$\mu ^{}(l)+p(r)\nu _\mu (l^{})+n(r^{})+\gamma (k),$$ (1.2) has a variable momentum transfer $`t`$ and one can get up to $`t=m_\mu ^2`$ at the maximum photon energy of about $`k100`$MeV, which is quite close to the pion pole. This amounts approximately to a four times larger sensitivity to $`g_P`$ in RMC than OMC. However, this increased sensitivity is upset by the very small partial branching ratio in hydrogen $`(10^8`$ for photons with $`k>60`$MeV) and one thus has to deal with large backgrounds. Precisely for this reason only very recently a first measurement of RMC on the proton has been published . The resulting number for $`g_P`$, which was obtained using a relativistic tree model including the $`\mathrm{\Delta }`$–isobar , came out significantly larger than expected from OMC, $`g_P^{\mathrm{RMC}}=12.35\pm 0.88\pm 0.38=1.46g_P^{\mathrm{CHPT}}`$. It should be noted that in this model the momentum dependence in $`G_P(t)`$ is solely given in terms of the pion pole and the induced pseudoscalar coupling is obtained as a multiplicative factor from direct comparison to the photon spectrum and the partial RMC branching ratio (for photon energies larger than 60 MeV). It was also argued in that the atomic and molecular physics related to the binding of the muon in singlet and triplet atomic $`\mu p`$ and ortho and para $`p\mu p`$ molecular states is sufficiently well under control. The TRIUMF result spurred a lot of theoretical activity. While radiative muon capture had already been calculated in phenomenological tree level models a long time ago, see e.g. , heavy baryon chiral perturbation theory was also used at tree level including dimension two operators and to one loop order . The resulting photon spectra are not very different from the ones obtained in the phenomenological models, the most striking feature being the smallness of the chiral loops , hinting towards a good convergence of the chiral expansion. At present, the puzzling result from the TRIUMF experiment remains unexplained. It is, however, a viable possibility that the discrepancy does not come from the strong interactions but rather is related to the distribution of the various spin states of the muonic atoms. It is therefore mandatory to sharpen the theoretical predictions for the strong as well as the non–strong physics entering the experimental analysis.<sup>#5</sup><sup>#5</sup>#5A recently proposed solution to the problem based on a novel term has been shown to be inconsistent with CVC in ref. . The corrected term–which in the chiral counting comes in at $`𝒪(p^3)`$–only gives a small contribution and is already contained in the RMC calculation of Ando and Min . Here, we wish to reanalyze RMC in the framework of the so–called small scale expansion , which allows to systematically include the $`\mathrm{\Delta }`$ resonance into the effective field theory. Although Ando and Min have already shown that the RMC process possesses a well behaved chiral expansion up to N<sup>2</sup>LO, it has been noted quite early that one should reanalyze RMC in a chiral effective field theory with explicit $`\mathrm{\Delta }`$ degrees of freedom<sup>#6</sup><sup>#6</sup>#6Phenomenological models have claimed for a long time that the $`\mathrm{\Delta }`$ contribution does not exceed 8% in the photon spectrum for photon energies above 60 MeV .. This is due to the fact that the $`\mathrm{\Delta }`$-resonance lies quite close to the nucleon and therefore, in a delta-free theory like HBChPT as used in , could lead to unnaturally large higher order contact interactions which would spoil the seemingly good chiral convergence. Stating the same concern in the language of (naive) dimensional analysis, it suggests the possibility of corrections of the order of 30% due to the small nucleon–delta mass splitting, $`m_\mu /(M_\mathrm{\Delta }M_N)3m_\mu /M_N`$. In the small scale expansion, the leading delta effects involving the large M1 $`\gamma N\mathrm{\Delta }`$ vertex already appear at second order $`ϵ^2`$ ($`ϵ`$ denotes a genuine small parameter, the pion mass, external momentum or the $`N\mathrm{\Delta }`$ mass splitting) and can therefore already be analyzed at tree level. The study presented here therefore constitutes a natural extension of the two previous analyses of RMC using chiral effective field theories<sup>#7</sup><sup>#7</sup>#7At the moment there exists no systematic scheme in chiral effective field theories to simultaneously include the effects of explicit vector mesons and other strong short range effects into the calculation. Given the small momentum transfer of RMC in the t-channel one expects, however, that these contributions can be correctly accounted for via $`𝒪(p^3)`$ contact interactions . A Born term analysis of vector meson contributions in the RMC process has been presented in .. Some preliminary results where reported in ref.. As we will show, the small scale expansion allows for a very transparent separation of the resonance and chiral pion effects. The manuscript is organized as follows. In section 2, we briefly review how the Standard Model at low energies is mapped onto a chiral effective field theory. The ingredients of this field theory, which uses pions, nucleons and the delta isobar as degrees of freedom, are discussed in section 3. Section 4 is concerned with ordinary and radiative muon capture on the proton. In the framework of the small scale expansion in OMC the leading order delta effects only come in at N<sup>2</sup>LO (i.e. $`𝒪(ϵ^3)`$), whereas for RMC one can already study the explicit influence of this resonance at NLO (i.e. $`𝒪(ϵ^2)`$). The status of the induced pseudoscalar form factor is reviewed and the analysis of the TRIUMF RMC experiment is critically assessed in section 5. Section 6 contains the summary and conclusions. Some more technical aspects of this work are relegated to the appendix. ## II From the Standard Model to the effective field theory In this section we briefly summarize how QCD coupled to the standard electroweak theory is mapped onto the pertinent effective field theory. For that, we start with the electroweak Lagrangian. For our purpose, we only need the coupling of the charged currents to the charged massive gauge bosons $`(W_\mu ^\pm `$) and the coupling of the massless photon ($`A_\mu `$) to the electromagnetic (em) current, $$_{\mathrm{int}}^{\mathrm{SM}}=\frac{g_2}{\sqrt{8}}\left\{W_\mu ^+(x)J_{\mathrm{ch}}^\mu +W_\mu ^{}(x)J_{\mathrm{ch}}^\mu \right\}g_1\mathrm{cos}\theta _WA_\mu (x)J_{\mathrm{em}}^\mu +_{\mathrm{ntl}\mathrm{wk}}$$ (2.1) where the last term refers to the non–leptonic weak interactions. $`g_1`$ and $`g_2`$ are the gauge couplings of the U(1)<sub>Y</sub> and the SU(2)<sub>L</sub> gauge groups and $`\theta _W`$ is the weak mixing (Weinberg) angle. In terms of the light quark and lepton fields the em and charged weak currents read $`J_{\mathrm{em}}^\mu `$ $`=`$ $`{\displaystyle \frac{2}{3}}\overline{u}\gamma ^\mu u{\displaystyle \frac{1}{3}}\overline{d}\gamma ^\mu d\overline{\mu }\gamma ^\mu \mu +\mathrm{}`$ (2.2) $`J_{\mathrm{ch}}^\mu `$ $`=`$ $`\overline{u}_w\gamma ^\mu \left(1\gamma _5\right)d_w+\overline{\nu }_\mu \gamma ^\mu \left(1\gamma _5\right)\mu +\mathrm{},`$ (2.3) with $`q_w=(u_w,d_w)`$ denoting the quark eigenstates of the weak interaction and the ellipsis denoting terms we do not need in what follows. For constructing the effective field theory, it is most convenient to consider the QCD Lagrangian coupled to these gauge fields treated as external local sources, which transform locally under chiral symmetry . With $`e=g_1\mathrm{cos}\theta _W`$ and $`W_\mu ^{}=𝒱_\mu ^{}(x)𝒜_\mu ^{}(x)`$, QCD coupled to these external sources takes the form $`_{\mathrm{QCD}}^{\mathrm{ext}.\mathrm{fields}}`$ $`=`$ $`_{\mathrm{QCD}}^0+\overline{q}\left[\mathbf{}(x)\mathbf{}(x)\gamma _5\right]q\overline{q}\left[𝐬(x)i𝐩(x)\right]q`$ (2.5) $`+\overline{q}\left[\mathit{}^{(0)}(x)\mathit{}^{(0)}(x)\gamma _5\right]q\overline{q}\left[s^{(0)}(x)ip^{(0)}(x)\right]q,`$ with $`q`$ the bi–spinor of the light quark fields in the strong interaction basis. For the case of RMC we are confronted with the following scenario of external sources<sup>#8</sup><sup>#8</sup>#8We are working in the limit of no isospin breaking, i.e. equal quark masses $`m_u=m_d`$ and no internal (virtual) photon effects.: $`𝐬(x)`$ $`=`$ $`0,`$ (2.6) $`s^{(0)}(x)`$ $`=`$ $`\widehat{m}I,`$ (2.7) $`𝐩(x)`$ $`=`$ $`p^{(0)}(x)=\mathrm{\hspace{0.33em}0},`$ (2.8) $`𝐯_\mu (x)`$ $`=`$ $`e{\displaystyle \frac{1}{2}}\tau ^3A_\mu (x)[{\displaystyle \frac{g_2V_{ud}}{\sqrt{8}}}({\displaystyle \frac{1}{2}}\tau ^1{\displaystyle \frac{i}{2}}\tau ^2)𝒱_\mu ^{}(x)+\mathrm{h}.\mathrm{c}.],`$ (2.9) $`v_\mu ^{(0)}(x)`$ $`=`$ $`e{\displaystyle \frac{1}{6}}IA_\mu (x),`$ (2.10) $`𝐚_\mu (x)`$ $`=`$ $`{\displaystyle \frac{g_2V_{ud}}{\sqrt{8}}}\left({\displaystyle \frac{1}{2}}\tau ^1{\displaystyle \frac{i}{2}}\tau ^2\right)𝒜_\mu ^{}(x)+\mathrm{h}.\mathrm{c}.,`$ (2.11) $`a_\mu ^{(0)}`$ $`=`$ $`0,`$ (2.12) with $`s(x),p(x),v_\mu (x)`$ and $`a_\mu (x)`$ scalar, pseudoscalar, vector and axial–vector fields, in order, in an obvious isospin (singlet and triplet) notation. $`V_{ud}`$ is the pertinent element of the CKM matrix. Note that the term $`_{\mathrm{QCD}}^0`$ in Eq.(2.5) is chirally symmetric. The explicit chiral symmetry breaking due to the current quark masses resides in the zeroth component of the external scalar source. Having specified the external field environment we now analyze the required strong matrix elements for calculating OMC and RMC. First we define quark vector and axial-vector currents $`V_\mu ^a`$ $`=`$ $`\overline{q}\gamma _\mu T^aq,`$ (2.13) $`A_\mu ^a`$ $`=`$ $`\overline{q}\gamma _\mu \gamma _5T^aq.`$ (2.14) Here, the $`T^a=\tau ^a/2`$ are the generators of SU(2). With these definitions one then specifies the strong matrix elements which have to be calculated in the effective theory $`a)`$ $`N|V_\mu ^a|N,`$ (2.15) $`b)`$ $`N|A_\mu ^a|N,`$ (2.16) $`c)`$ $`N|𝒯V_\mu ^aV_\nu ^b|N,`$ (2.17) $`d)`$ $`N|𝒯V_\mu ^aA_\nu ^b|N,`$ (2.18) with all possible combinations of the external fields of Eqs.(2.6-2.12) and $`𝒯`$ denotes the conventional time–ordering operator. To proceed, we now have to specify the effective Lagrangian which will be used to calculate these matrix elements. ## III Chiral Lagrangians In this section, we briefly review the chiral Lagrangian underlying our calculation. While the meson part is standard, with the external momenta and the pion (light quark) mass counted as small parameters, for the meson–baryon system we include the nucleons as well as the $`\mathrm{\Delta }(1232)`$ resonance. To systematically account for the effects of the latter, the $`N\mathrm{\Delta }`$ mass splitting is considered as an additional small parameter. This is routed in phenomenology and the large $`N_c`$ limit of QCD, but not in the strict chiral limit, where the $`\mathrm{\Delta }`$ decouples. These three small parameters are collectively denoted by $`ϵ`$. ### A Meson chiral perturbation theory At low energies the chiral symmetry of QCD is spontaneously broken to the vectorial subgroup $`SU(2)_V`$: $`SU(2)_L\times SU(2)_RSU(2)_V`$. The strictures of the spontaneous and the explicit chiral symmetry breaking can be explored in terms of an effective field theory, chiral perturbation theory (ChPT). As a tool, one works with an effective Lagrangian, which consists of a string of terms with increasing dimension. For our purpose, we only need the first term in this expansion, the non-linear $`\sigma `$–model chirally coupled to the external sources, $`_{\pi \pi }^{(2)}`$ $`=`$ $`{\displaystyle \frac{F_0^2}{4}}\mathrm{Tr}\left[_\mu U^{}^\mu U+\chi ^{}U+\chi U^{}\right]`$ (3.1) with $`U(x)`$ $`=`$ $`\mathrm{exp}\left\{{\displaystyle \frac{i}{F_0}}\tau \pi (x)\right\},`$ (3.2) $`_\mu U`$ $`=`$ $`_\mu Ui(𝐯_\mu +𝐚_\mu )U+iU(𝐯_\mu 𝐚_\mu ),`$ (3.3) $`\chi `$ $`=`$ $`2B_0\left(𝐬+s^{(0)}+i𝐩+ip^{(0)}\right).`$ (3.4) The pions, which are nothing but integration variables, are collected in the matrix–valued field $`U(x)`$ and the explicit symmetry breaking due to the light quark masses is hidden in $`\chi `$ via the zeroth component of the scalar source. $`F_0`$ is the (weak) pion decay constant (in the chiral limit) and $`B_0`$ is related to the scalar quark condensate. We work in the standard scenario with $`B_0F_0`$. This specifies completely the meson part of the effective Lagrangian. ### B Including baryons: The small scale expansion The nucleon–delta–pion system chirally coupled to the external fields can also be represented by a Lagrangian, which decomposes into a string of terms with increasing dimension. We work here in the heavy mass formulation, in which the nucleon and the delta are essentially considered as heavy, static sources. This allows to shuffle the baryon mass ($`M_B`$) dependence into a string of $`1/M_B`$ suppressed vertices and gives rise to a consistent power counting . Denoting by $`N`$ the (heavy) nucleon isodoublet field and by $`T_\mu ^i`$ the Rarita–Schwinger representation of the heavy spin-3/2 field, the lowest order terms read $`_{\pi N}^{(1)}`$ $`=`$ $`\overline{N}\left[ivD+\dot{g}_ASu\right]N,`$ (3.5) $`_{\pi \mathrm{\Delta }}^{(1)}`$ $`=`$ $`\overline{T}_i^\mu \left[ivD^{ij}\delta ^{ij}\mathrm{\Delta }_0+\mathrm{}\right]g_{\mu \nu }T_j^\nu ,`$ (3.6) $`_{\pi N\mathrm{\Delta }}^{(1)}`$ $`=`$ $`\dot{g}_{\pi N\mathrm{\Delta }}\left\{\overline{T}_i^\mu g_{\mu \alpha }w_i^\alpha N+\overline{N}w_i^\alpha g_{\alpha \mu }T_i^\mu \right\},`$ (3.7) with $`\mathrm{\Delta }_0=M_\mathrm{\Delta }M_0(M_0)`$ being the nucleon-delta mass splitting (nucleon mass) in the chiral limit (to the order we are working, we can set $`\mathrm{\Delta }_0=\mathrm{\Delta }`$). All other quantities $`Q`$ in the chiral limit are denoted as $`\dot{Q}`$. Furthermore, $`v_\mu `$ and $`S_\mu `$ are the four–velocity and the spin–vector of the heavy nucleon. The various chiral covariant derivatives, chiral connections and vielbeins are (for details, we refer to and ) $`D_\mu N`$ $`=`$ $`(_\mu +\mathrm{\Gamma }_\mu iv_\mu ^{(s)})N,`$ (3.8) $`v_\mu ^{(s)}`$ $`=`$ $`3v_\mu ^{(0)},`$ (3.9) $`\mathrm{\Gamma }_\mu `$ $`=`$ $`{\displaystyle \frac{1}{2}}[u^{},_\mu u]{\displaystyle \frac{i}{2}}u^{}(𝐯_\mu +𝐚_\mu )u{\displaystyle \frac{i}{2}}u(𝐯_\mu 𝐚_\mu )u^{}\tau ^i\mathrm{\Gamma }_\mu ^i,`$ (3.10) $`u_\mu `$ $`=`$ $`iu^{}_\mu Uu^{}\tau ^iw_\mu ^i,`$ (3.11) $`D_\mu ^{ij}T_\nu ^j`$ $`=`$ $`\left(_\mu \delta ^{ij}+C_\mu ^{ij}\right)T_\nu ^j,`$ (3.12) $`C_\mu ^{ij}`$ $`=`$ $`\delta ^{ij}\left(\mathrm{\Gamma }_\mu iv_\mu ^{(s)}\right)2iϵ^{ijk}\mathrm{\Gamma }_\mu ^k,`$ (3.13) where $`i,j,k`$ are isospin indices. At second order in the small scale expansion, we need the following terms $`_{\pi N}^{(2)}`$ $`=`$ $`{\displaystyle \frac{1}{2M_0}}\overline{N}\{(vD)^2D^2ig_A(SDvu+vuSD)`$ (3.15) $`{\displaystyle \frac{i}{2}}[S^\mu ,S^\nu ]_{}[(1+\dot{\kappa }_v)f_{\mu \nu }^++2(1+\dot{\kappa }_s)v_{\mu \nu }^{(s)}]+\mathrm{}\}N,`$ $`_{\pi N\mathrm{\Delta }}^{(2)}`$ $`=`$ $`\overline{T}_i^\mu {\displaystyle \frac{1}{2M_0}}\left[b_1if_{+\mu \nu }^iS^\nu +\mathrm{}\right]N+\mathrm{h}.\mathrm{c}.`$ (3.16) with $`f_{\mu \nu }^\pm `$ $`=`$ $`u^{}F_{\mu \nu }^Ru\pm uF_{\mu \nu }^Lu^{}\tau ^if_{\pm \mu \nu }^i,`$ (3.17) $`F_{\mu \nu }^X`$ $`=`$ $`_\mu F_\nu ^X_\nu F_\mu ^Xi[F_\mu ^X,F_\nu ^X];X=L,R,`$ (3.18) $`F_\mu ^R`$ $`=`$ $`𝐯_\mu +𝐚_\mu ,F_\mu ^L=𝐯_\mu 𝐚_\mu ,`$ (3.19) $`v_{\mu \nu }^{(s)}`$ $`=`$ $`_\mu v_\nu ^{(s)}_\nu v_\mu ^{(s)}.`$ (3.20) Note that in $`_{\pi N}^{(2)}`$ only two low–energy constants (LECs) appear, which we have expressed in terms of the isoscalar and isovector anomalous magnetic moments, $`\kappa _s`$ and $`\kappa _v`$, of the nucleon, respectively. The LEC $`b_1`$ is fixed from neutral pion photoproduction at threshold, $`b_1=12.0`$. The $`𝒪(ϵ^2)`$ Lagrangians given so far are sufficient to calculate the leading order $`\mathrm{\Delta }`$(1232) effects in RMC. However, as mentioned above, in OMC the leading $`\mathrm{\Delta }`$(1232) related effects only occur at N<sup>2</sup>LO—specifically as a $`𝒪(ϵ^3)`$ contribution to the form factors of the nucleon. In SSE these terms have already been calculated in and we refer the reader interested in the details of the third order Lagrangian required for such a calculation to ref.. We now have specified the effective Lagrangian needed to work out OMC and RMC to second order in the small scale expansion. The pertinent Feynman rules to perform these calculations are collected in appendix A. ## IV Muon capture This section is concerned with the theoretical description of OMC and RMC. To keep the manuscript self–contained, we give all formulae necessary to calculate the capture rates. ### A Leptonic matrix elements Consider first the purely leptonic part of the muon capture reaction. For OMC and RMC, we need the following two leptonic matrix elements: $`\nu _\mu |J_\mu ^+|\mu `$ $`=`$ $`i{\displaystyle \frac{g_2}{\sqrt{8}}}\overline{\nu }_\mu (l^{})\gamma _\mu \left(1\gamma _5\right)\mu (l),`$ (4.1) $`\nu _\mu \gamma |J_\mu ^+|\mu `$ $`=`$ $`i{\displaystyle \frac{g_2}{\sqrt{8}}}{\displaystyle \frac{e}{2lk}}\overline{\nu }_\mu (l^{})\gamma _\mu \left(1\gamma _5\right)\left(2ϵ^{}l\mathit{}\mathit{ϵ̸}^{}\right)\mu (l),`$ (4.2) with $`ϵ_\mu ^{}`$ the polarization vector of the photon. The corresponding four–momenta have already been defined in Eq.(1.1) and Eq.(1.2), respectively. ### B Hadronic matrix elements For OMC one only needs to calculate the vector and axial-vector transition matrix elements $`n|V_\mu ^{}|p,n|A_\mu ^{}|p`$ of Eqs.(2.15,2.16). In the limit of exact isospin symmetry these matrix-elements constitute the off-diagonal components of the isovector nucleon vector and axial-vector currents which are typically parameterized in terms of four form factors (e.g. ref.): the isovector vector form factors $`F_{1,2}^v`$, as well as the axial and the induced pseudoscalar form factor, $`G_A(q^2)`$ and $`G_P(q^2)`$ respectively. In terms of these, the relativistic strong matrix elements defined in Eqs.(2.15,2.16) (which corresponds to the vector and axial correlators, shown in the first row of fig.1) are given by: $`n|V_\mu ^{}|p`$ $`=`$ $`i{\displaystyle \frac{g_2V_{ud}}{\sqrt{8}}}\overline{n}(p_2)\left[F_1^v(q^2)\gamma _\mu +{\displaystyle \frac{i}{2M_N}}F_2^v(q^2)\sigma _{\mu \nu }q^\nu \right]p(p_1),`$ (4.3) $`n|A_\mu ^{}|p`$ $`=`$ $`i{\displaystyle \frac{g_2V_{ud}}{\sqrt{8}}}\overline{n}(p_2)\left[G_A(q^2)\gamma _\mu \gamma _5+{\displaystyle \frac{G_P(q^2)}{2M_N}}q_\mu \gamma _5\right]p(p_1).`$ (4.4) where $`q^2=(p_2p_1)^2`$ is the invariant momentum transfer squared. The Dirac $`F_1^{(v)}`$ and Pauli $`F_2^{(v)}`$ form factor are subject to the normalizations $`F_1^{(v)}(0)=1,F_2^{(v)}(0)=\kappa _v,`$ (4.5) with $`\kappa _v=3.71`$ the isovector nucleon anomalous magnetic moment. In the axial matrix element, Eq.(4.4) one has assumed the absence of second class currents. The electromagnetic form factor are rather well known. The current situation concerning their theoretical understanding and experimental knowledge can be found e.g. in . Muon captures therefore provide us with the opportunity to study the weak axial structure of a nucleon. While $`G_A(q^2)`$ can be extracted from (anti)neutrino–proton scattering or charged pion electroproduction data, $`G_P(q^2)`$ is harder to pin down and in fact constitutes the least known nucleon form factor. In Fig.7 we present the “world data” for $`G_P(q^2)`$. These four form factors have already been calculated to $`𝒪(ϵ^3)`$ in . One finds<sup>#9</sup><sup>#9</sup>#9Note that due to the small momentum transfer $`|q^2|<0.01`$ GeV<sup>2</sup> in OMC/RMC it is sufficient to work in the radius approximation of the form factors, i.e. to truncate the $`q^2`$-dependence after the first term. The full $`𝒪(ϵ^3)`$ momentum dependence for $`q^2<0.2`$ GeV<sup>2</sup> can be found in . $`F_i^{(v)}(q^2)`$ $`=`$ $`F_i^{(v)}(0)\left[1+{\displaystyle \frac{1}{6}}(r_i^v)^2q^2+𝒪(q^4)\right]`$ (4.6) $`G_A(q^2)`$ $`=`$ $`g_A\left[1+{\displaystyle \frac{1}{6}}(r_A)^2q^2+𝒪(q^4)\right]`$ (4.7) $`G_P(q^2)`$ $`=`$ $`{\displaystyle \frac{4M_Ng_{\pi NN}F_\pi }{m_\pi ^2q^2}}{\displaystyle \frac{2}{3}}g_AM_N^2r_A^2.`$ (4.8) where we have systematically shifted all appearing quantities to their physical values. $`r_{1,2}^v`$ and $`r_A`$ are the isovector Dirac and Pauli radius and the axial radius respectively. They receive contributions from chiral loops and counter-terms, except for $`r_2^v`$ which is free of any LEC. Detailed results can be found in . In Fig.7 the difference between the usual pion-pole parameterization for $`G_P(q^2)`$ and the full chiral structure of the form factor, Eq.4.8 is displayed. Note that this chiral structure is not affected by $`\mathrm{\Delta }`$(1232). In the kinematical region of RMC, which mainly lies to the “left” of the OMC point in Fig.7 the structure effect in $`G_P(q^2)`$ proportional to the axial radius $`r_A`$ is expected to play only a small role. Certainly, the present experimental uncertainties both in OMC and in RMC are too large to distinguish between the two curves, but new efforts are under way <sup>#10</sup><sup>#10</sup>#10 Let us briefly emphasize that there exists another window on $`G_P(q^2)`$—pion electroproduction. So far there has only been one experiment that took up the challenge, with the results shown in Fig.7. In this kinematical regime the structure proportional to $`r_A`$ produces the biggest effect and a new dedicated experiment should be able to identify it — thereby enhancing our knowledge of this poorly known form factor considerably! In fact, at the Mainz Microtron MAMI-B a dedicated experiment has been proposed to measure the axial and the induced pseudoscalar form factors by means of charged pion electroproduction at low momentum transfer .. Usually the non relativistic reduction of Eqs.(4.3,4.4) is done in the Breit frame. Here we give the non relativistic strong matrix elements in the rest frame of the proton where our calculation is done: $`n|V_\mu ^{}|p`$ $`=`$ $`i{\displaystyle \frac{g_2V_{ud}}{\sqrt{8}}}𝒩_2\overline{n}_v(p_2)\{({\displaystyle \frac{2M_N}{E_2+M_N}}F_1^{(v)}(q^2){\displaystyle \frac{E_2M_N}{E_2+M_N}}F_2^{(v)}(q^2))v_\mu `$ (4.11) $`+\left[{\displaystyle \frac{1}{E_2+M_N}}\left(F_1^{(v)}(q^2)+F_2^{(v)}(q^2)\right){\displaystyle \frac{1}{2M_N}}F_2^{(v)}(q^2)\right]q_\mu `$ $`+{\displaystyle \frac{2}{E_2+M_N}}[S_\mu ,Sq](F_1^{(v)}(q^2)+F_2^{(v)}(q^2))\}p_v(0),`$ $`n|A_\mu ^{}|p`$ $`=`$ $`i{\displaystyle \frac{g_2V_{ud}}{\sqrt{8}}}𝒩_2\overline{n}_v(p_2)\{G_A(q^2)[2S_\mu {\displaystyle \frac{2Sqv_\mu }{E_2+M_N}}]`$ (4.13) $`+RG_P(q^2){\displaystyle \frac{Sqq_\mu }{M_N\left(E_2+M_N\right)}}\}p_v(0),`$ where $`𝒩_2`$ is the usual normalization factor of the neutron wave function, $`𝒩_2=\sqrt{\frac{E_2+M_N}{2M_N}}`$ and $`E_2`$ is the neutron energy. One gets immediately the results to $`𝒪(ϵ^2)`$ by replacing the form factors by their values at $`q^2=0`$ We now turn to the vector–vector (VV) and vector–axial (VA) correlator which we only need to $`𝒪(ϵ^2)`$ in SSE. Working in the Coulomb gauge $`ϵ^{}v=0`$ for the photon and making use of the transversality condition $`ϵ^{}k=0`$, we find (the pertinent Feynman diagrams for the VV correlator are shown in the second row in fig. 1 and the ones for VA in figs. 2,3) $`n|𝒯Vϵ^{}V_\mu ^{}|p^{(2)}`$ $`=`$ $`i{\displaystyle \frac{g_2V_{ud}e}{\sqrt{8}}}\overline{n}_v(r^{})\{{\displaystyle \frac{1+\kappa _v}{M_N}}[S_\mu ,Sϵ^{}]{\displaystyle \frac{1}{2M_N}}ϵ_\mu ^{}`$ (4.15) $`+{\displaystyle \frac{1}{M_N\omega }}v_\mu [(1+\kappa _v)[Sϵ^{},Sk]ϵ^{}r]+𝒪(1/M_N^2)\}p_v(r),`$ $`n|𝒯Vϵ^{}A_\mu ^{}|p^{(2)}`$ $`=`$ $`i{\displaystyle \frac{g_2V_{ud}e}{\sqrt{8}}}\overline{n}_v(r^{})\times `$ (4.23) $`\{2Rg_A{\displaystyle \frac{S(r^{}r)}{(r^{}r)^2m_\pi ^2}}\times [{\displaystyle \frac{2ϵ^{}(ll^{})(ll^{})_\mu }{(ll^{})^2m_\pi ^2}}ϵ_\mu ^{}]`$ $`R{\displaystyle \frac{g_A}{M_N}}{\displaystyle \frac{\left(vr^{}vr\right)S(r+r^{})}{(r^{}r)^2m_\pi ^2}}\times \left[{\displaystyle \frac{2ϵ^{}(ll^{})(ll^{})_\mu }{(ll^{})^2m_\pi ^2}}ϵ_\mu ^{}\right]`$ $`2Rg_A\left[1+{\displaystyle \frac{vlvl^{}}{2M_N}}\right]{\displaystyle \frac{Sϵ^{}(ll^{})_\mu }{(ll^{})^2m_\pi ^2}}+{\displaystyle \frac{g_A}{M_N}}Sϵ^{}v_\mu `$ $`+{\displaystyle \frac{g_A}{M_N}}[{\displaystyle \frac{(2+\kappa _s+\kappa _v)S^\alpha [Sϵ^{},Sk]}{\omega }}+{\displaystyle \frac{(\kappa _v\kappa _s)[Sϵ^{},Sk]S^\alpha }{\omega }}`$ $`{\displaystyle \frac{2S^\alpha ϵ^{}r}{\omega }}]\times [g_{\mu \alpha }R{\displaystyle \frac{(ll^{})_\alpha (ll^{})_\mu }{(ll^{})^2m_\pi ^2}}]`$ $`+{\displaystyle \frac{g_{\pi N\mathrm{\Delta }}b_1}{3M_N}}[{\displaystyle \frac{2\mathrm{\Delta }[k^\alpha Sϵ^{}\omega v^\alpha Sϵ^{}ϵ^\alpha Sk]}{\mathrm{\Delta }^2\omega ^2}}{\displaystyle \frac{4S^\alpha [Sϵ^{},Sk]}{3(\mathrm{\Delta }+\omega )}}`$ $`+{\displaystyle \frac{4[Sϵ^{},Sk]S^\alpha }{3(\mathrm{\Delta }\omega )}}]\times [g_{\mu \alpha }{\displaystyle \frac{(ll^{})_\alpha (ll^{})_\mu }{(ll^{})^2m_\pi ^2}}]+𝒪(1/M_N^2)\}p_v(r),`$ with $`\omega =vk`$ and $`R=1`$ in QCD. We have introduced this factor multiplying the Born term contributions proportional to the induced pseudoscalar form factor for the later discussion. One can easily check from the continuity equations satisfied by the correlators (which for example relates the vector–axial correlator to the axial one, see ) that gauge invariance is satisfied in the above equations. We note again that here we only give the results to $`𝒪(ϵ^2)`$ as this is sufficient to study the leading order $`\mathrm{\Delta }`$(1232) effects in RMC. We also want to point out that the vector–vector correlator is free of delta effects to this order $`𝒪(ϵ^2)`$, i.e. the leading $`\mathrm{\Delta }`$(1232) effect only appears in the vector–axial correlator, cf. fig 3. ### C Ordinary Muon Capture Surprisingly, no full analysis of OMC exists in the literature for the framework of chiral effective field theories. All previous analyses stopped at the level of writing down the relevant matrix-elements/form factors (here Eqs.(4.3,4.4)), but no discussion of the implications for the lifetime of OMC was given, which can be written down in a closed analytic form. In this section we are going to fill this gap. We work in the Fermi approximation of a static $`W_\mu ^{}`$ field, i.e. the gauge boson propagator is reduced to a point interaction (since the typical momenta involved are much smaller than the $`W`$ mass), $`_{\mu ^{}p\nu _\mu n}`$ $`=`$ $`^{\mathrm{OMC}}=\nu _\mu |W_\mu ^+|\mu i{\displaystyle \frac{g^{\mu \nu }}{M_W^2}}\left[n|V_\nu ^{}|pn|A_\nu ^{}|p\right].`$ (4.24) #### 1 Spin-averaged OMC Introducing the Fermi constant $`G_F`$ via $`G_F=g_2^2\sqrt{2}/(8M_W^2)`$, we define the square of the spin-averaged invariant matrix element to be $`{\displaystyle \frac{1}{4}}{\displaystyle \underset{\sigma \sigma ^{}ss^{}}{}}|^{\mathrm{OMC}}|^2`$ $`=`$ $`{\displaystyle \frac{G_F^2V_{ud}^2}{2}}L_{\mu \nu }^{(a)}H_{(a)}^{\mu \nu }.`$ (4.25) With the normalizations $`{\displaystyle \underset{s}{}}\mu (l,s)\overline{\mu }(l,s)={\displaystyle \frac{\mathit{}+m_\mu ^{}}{2m_\mu ^{}}},`$ $`{\displaystyle \underset{s}{}}\nu (l^{},s)\overline{\nu }(l^{},s)=\mathit{}^{},`$ (4.26) $`{\displaystyle \underset{\sigma }{}}p_v(r,\sigma )\overline{p}_v(r,\sigma )=P_v^+\left(1+{\displaystyle \frac{vr}{2M_N}}\right),`$ $`{\displaystyle \underset{\sigma }{}}n_v(r^{},\sigma )\overline{n}_v(r^{},\sigma )=P_v^+\left(1+{\displaystyle \frac{vr^{}}{2M_N}}\right),`$ (4.27) one then obtains the symmetric tensors $`L_{\mu \nu }^{(a)}`$ $`=`$ $`{\displaystyle \frac{2}{m_\mu ^{}}}\left\{l_\mu l_\nu ^{}g_{\mu \nu }ll^{}+l_\mu ^{}l_\nu +iϵ_{\mu \alpha \nu \beta }l^\alpha l^\beta \right\},`$ (4.28) $`H_{\mu \nu }^{(a)}`$ $`=`$ $`av_\mu v_\nu +b\left(v_\mu v_\nu g_{\mu \nu }\right)+cr_\mu ^{}r_\nu ^{}+d\left(v_\mu r_\nu ^{}+r_\mu ^{}v_\nu \right)+eiϵ_{\mu \nu }^{\alpha \beta }r_\alpha ^{}v_\beta .`$ (4.29) Note that we have evaluated the tensors for the special kinematic condition of both the proton and the muon being at rest<sup>#11</sup><sup>#11</sup>#11This approximation is well-justified due to the low binding energy $`E2.5`$keV of an s-wave muonic atom as compared to the muon mass., i.e. $`l_\mu =(m_\mu ^{},0,0,0),r_\mu =(0,0,0,0)`$. $`a,b,c,d,e`$ are coefficients which depend on the four form factors Eqs.(4.6, 4.7). For illustration we give them to $`𝒪(ϵ^2)`$ $`a=1,b=g_A^2,c=g_A^2{\displaystyle \frac{\stackrel{}{r}^{\mathrm{\hspace{0.17em}2}}+2m_\pi ^2}{(\stackrel{}{r}^{\mathrm{\hspace{0.17em}2}}+m_\pi ^2)^2}},d={\displaystyle \frac{1+g_A^2}{2M_N}},e={\displaystyle \frac{g_A(1+\kappa _v)}{M_N}}.`$ (4.30) We now assume that the initial muon-proton system constitutes the ground-state of a bound system. We therefore replace the plane-wave wavefunction of the muon used so far in the calculation by the 1s Bohr-wavefunction $`\mathrm{\Phi }(x)_{1s}`$ of a muonic atom. We note that to $`𝒪(ϵ^2)`$ we are not sensitive to the extended structure of the proton – any dependence of the capture process on the electric/magnetic radius of the nucleon will only occur at $`𝒪(ϵ^3)`$ . As we will see below, the terms of $`𝒪(ϵ^3)`$ in OMC only represent a small N<sup>2</sup>LO correction. For simplicity we therefore approximate the muonic atom wave-function dependence by its value at the origin even at $`𝒪(ϵ^3)`$, which effectively constitutes an upper bound for the $`𝒪(ϵ^3)`$ correction to the total width. In a systematic $`𝒪(ϵ^3)`$ analysis one would of course have to calculate the overlap between the Bohr-wavefunction and the respective proton-neutron transition form factors. Confining ourselves to $`x=0`$ we use $$\mathrm{\Phi }(0)_{1s}=\frac{\alpha ^{3/2}\mu ^{3/2}}{\sqrt{\pi }},$$ (4.31) with the reduced mass $`\mu =M_Nm_\mu ^{}/(M_N+m_\mu ^{})`$ and $`\alpha =e^2/4\pi \mathrm{}c`$ in the Heaviside-Lorentz convention. We can therefore calculate the spin-averaged rate of ordinary muon capture via $`\mathrm{\Gamma }_{\mathrm{OMC}}`$ $`=`$ $`|\mathrm{\Phi }(0)_{1s}|^2{\displaystyle \frac{d^3r^{}}{(2\pi )^3J_n}\frac{d^3l^{}}{(2\pi )^3J_\nu }\left(2\pi \right)^4\delta ^4\left(r+lr^{}l^{}\right)\frac{1}{4}\underset{\sigma \sigma ^{}ss^{}}{}|^{\mathrm{OMC}}|^2}`$ (4.32) with the normalization factors $`J_\nu =2E_\nu ,J_n=1+vr^{}/M_N`$. Evaluating all expressions at $`𝒪(ϵ^3)`$ accuracy one obtains $`\mathrm{\Gamma }_{\mathrm{OMC}}^{\mathrm{SSE}}`$ $`=`$ $`\left(\mathrm{\hspace{0.17em}247.0}\mathrm{\hspace{0.17em}61.6}4.0+𝒪(1/M_N^3)\right)\times \mathrm{s}^1`$ (4.33) $`=`$ $`181.5\times \mathrm{s}^1`$ (4.34) Here we have used the coupling constants, masses and radii as given in table I. $`r_1^{(v)}`$ and $`r_A`$ corresponds to their empirical values since the value of the counter-terms which enter in their expressions to $`𝒪(ϵ^3)`$ has been set to reproduce them. The value of $`(r_2^{(v)})_{\mathrm{SSE}}^2=0.61`$ fm<sup>2</sup> is given counter-term free by the $`𝒪(ϵ^3)`$ SSE calculation. For comparison we also give the spin averaged OMC capture rate to $`𝒪(p^3)`$ in HBCHPT without explicit spin 3/2 degrees of freedom: $`\mathrm{\Gamma }_{\mathrm{OMC}}^{HBChPT}`$ $`=`$ $`\left(\mathrm{\hspace{0.17em}247.0}\mathrm{\hspace{0.17em}61.6}3.8+𝒪(1/M_N^3)\right)\times \mathrm{s}^1`$ (4.35) $`=`$ $`181.7\times \mathrm{s}^1`$ (4.36) The only change occurs at N<sup>2</sup>LO via $`(r_2^{(v)})_{\mathrm{HBChPT}}^2=0.52`$ fm<sup>2</sup>, showing only a very weak dependence on the exact value of the isovector Pauli radius, which has the “exact” value $`(r_2^{(v)})_{exp.}^2=0.80`$ fm<sup>2</sup>. We note that in Eqs.(4.33) the $`𝒪(ϵ^2)`$ SSE contribution amounts to a correction of 25% of the leading term, while the $`𝒪(ϵ^3)`$ one is two orders of magnitude smaller, indicating that muon capture on a proton really constitutes a system with an extremely well behaved chiral expansion. Finally we discuss the case of no explicit chiral symmetry breaking (i.e. in the chiral limit $`m_\pi =0`$). One expects the spin-averaged capture rate to behave as $`\mathrm{\Gamma }_{\mathrm{OMC}}^\chi `$ $`=`$ $`\left(\mathrm{\hspace{0.17em}214}\mathrm{\hspace{0.17em}46}+𝒪(1/M_N^2)\right)\times \mathrm{s}^1`$ (4.37) $`=`$ $`168\times \mathrm{s}^1.`$ (4.38) which only represents a small finite shift compared to the value for finite quark masses. The leading infrared singularities only occur at N<sup>2</sup>LO due to the long range nature of the chiral pion cloud in the isovector form factor radii . The reason for the nice stability of perturbative calculations for OMC in the physical world of small finite quark masses is of course the fact that contributions of order $`n`$ are suppressed by $`\left(m_i/\mathrm{\Lambda }_\chi \right)^{n1}`$, with $`i=\pi ,\mu `$ and $`\mathrm{\Lambda }_\chi M_N1`$ GeV. Note that one could also retain the higher order kinematical corrections starting at order $`1/M_N^3`$. To be more precise, this refers to the energy–momentum relation between the various particles, the mass term appearing in the various projection operators and the phase space. In that case, only the various correlators (A,V,AV,VV) are truncated at order $`1/M_N`$. Using this approximation, the total rate is $`\mathrm{\Gamma }_{\mathrm{OMC}}=193\mathrm{s}^1`$, not much different from the one given in Eq.(4.33). We remark already at this point that in the case of RMC, it is mandatory to retain these higher order terms if one stays at a low order in the small scale expansion as done here. #### 2 Hyperfine effect in OMC The spin-averaged OMC scenario presented in the previous section is only of theoretical interest. In nature the weak-interactions shows a very strong spin-dependence, which leads to quite different decay-rates depending on whether the captured 1s muon forms a singlet or a triplet spin-state with its proton , where the singlet is the usual state $`(1/\sqrt{2})(|,|,)`$ in terms of the muon and proton spins and the triplet accordingly. Although the hyperfine splitting between the two levels “only” amounts to 0.04 eV, the occupation numbers of the levels due to thermal and collision induced processes tend to be far from statistical equilibrium. In order to make any contact with experiment, the singlet/triplet rates need to be calculated separately. These various spin states are obtained by using the following projection operators for the muon, $`\mu (l,1/2)\overline{\mu }(l,1/2)`$ $`=`$ $`{\displaystyle \frac{1}{2}}\left(1+\gamma _5\overline{)𝑠}\right){\displaystyle \frac{\overline{)𝑙}+m_\mu }{2m_\mu }},`$ (4.39) $`\mu (l,\pm 1/2)\overline{\mu }(l,1/2)`$ $`=`$ $`{\displaystyle \frac{\overline{)𝑙}+m_\mu }{2m_\mu (E_\mu +m_\mu )}}\gamma _5\gamma _0{\displaystyle \frac{1}{2}}(\gamma _1\pm i\gamma _2)\gamma _0(\overline{)𝑙}+m_\mu ),`$ (4.40) and similarly for the proton, $`u(l,1/2)\overline{u}(l,1/2)`$ $`=`$ $`{\displaystyle \frac{1}{2}}\left(1+\gamma _5\overline{)𝑠}\right){\displaystyle \frac{1}{2}}(1+\overline{)𝑣})\left(1+{\displaystyle \frac{vr}{2M_N}}\right),`$ (4.41) $`u(l,\pm 1/2)\overline{u}(l,1/2)`$ $`=`$ $`{\displaystyle \frac{1}{2}}(1+\overline{)𝑣}){\displaystyle \frac{\overline{)𝑟}+2M_N}{2M_N(E_N+M_N)}}\gamma _5\gamma _0{\displaystyle \frac{1}{2}}(\gamma _1\pm i\gamma _2)\gamma _0(\overline{)𝑟}+2M_N){\displaystyle \frac{1}{2}}(1+\overline{)𝑣}).`$ (4.42) For the total capture rates of singlet and the triplet states in the muonic atom, we then find the following decomposition into leading, next–to–leading order pieces and next–to–next–to–leading order pieces to $`𝒪(ϵ^3)`$ in SSE<sup>#12</sup><sup>#12</sup>#12For the remaining part of the discussion on OMC we only give the values calculated in to $`𝒪(ϵ^3)`$ in the small scale expansion. The corresponding values to $`𝒪(p^3)`$ in HBChPT are very similar. For OMC the two effective field theories only differ in the N<sup>2</sup>LO contribution of the isovector Pauli radius as discussed in the previous section.: $`\mathrm{\Gamma }_{\mathrm{OMC}}^{\mathrm{sing}}`$ $`=`$ $`(957245\mathrm{GeV}/M_N+(30.4\mathrm{GeV}/M_N^243.17)+𝒪(1/M_N^3))\times \mathrm{s}^1=687.4\times \mathrm{s}^1,`$ (4.43) $`\mathrm{\Gamma }_{\mathrm{OMC}}^{\mathrm{trip}}`$ $`=`$ $`(10.3+4.72\mathrm{GeV}/M_N(1.22\mathrm{GeV}/M_N^2+1.00)+𝒪(1/M_N^3))\times \mathrm{s}^1=12.9\times \mathrm{s}^1,`$ (4.44) displaying the dramatic spin-dependence due to the V-A structure of the weak interaction in the Standard Model. These numbers correspond to a value of $`g_{\pi NN}`$ as given in table I and includes the pion pole corrections, Eq.(4.8). Since $`G_P`$ contributes negatively to the singlet rate a larger value of $`g_{\pi NN}`$ leads to a smaller value for the rate: $`\mathrm{\Gamma }_{\mathrm{OMC}}^{\mathrm{sing}}=681.9\times \mathrm{s}^1`$ for $`g_{\pi NN}=13.4`$. Similarly, neglecting the pion pole corrections leads to $`\mathrm{\Gamma }_{\mathrm{OMC}}^{\mathrm{sing}}=676.1\times \mathrm{s}^1`$. In Eq.(4.44) we have split the $`𝒪(ϵ^3)`$ term (third and fourth terms in parenthesis) into the contribution from the $`1/M^2`$ corrections and the pure $`𝒪(ϵ^3)`$ terms stemming from the various radii which lead to the $`q^2`$ dependence of the form factors, Eq.(4.8). It is very interesting to note that these two contributions more or less cancels themselves in the case of the singlet term. It is thus extremely important to perform a consistent chiral expansion. In a relativistic Born model one obtains the simple nice formula<sup>#13</sup><sup>#13</sup>#13We note that we have different relative signs compared to ref. to translate into our notation. for $`\mathrm{\Gamma }_{\mathrm{OMC}}^{\mathrm{sing}}`$ : $`\mathrm{\Gamma }_{\mathrm{OMC}}^{\mathrm{sing}}`$ $``$ $`\left(6.236F_1^v(q_0^2)+0.5513F_2^v(q_0^2)+16.44G_A(q_0^2)0.2834G_P(q_0^2)\right)^2`$ (4.45) $``$ $`683\times \mathrm{s}^1`$ (4.46) where $`q_0^2=0.88m_\mu ^2`$ is the momentum transfer and the remaining parameters are taken from table I. The smaller value quoted in , comes from the fact that in the sixties $`g_A`$ was somewhat smaller and $`r_A`$ somewhat bigger. This formula though very appealing should not be used, being in contradiction with the modern viewpoint of power counting. The good agreement with the SSE result is purely accidental. The problem arises from the fact that $`\mathrm{\Gamma }_{\mathrm{OMC}}^{\mathrm{sing}}`$ is a rather sensitive quantity as can be seen for example in Eq.(4.45). Indeed the terms proportional to $`F_2^v(q_0^2)`$ and $`G_P(q_0^2)`$ are of the same order of magnitude but have different signs so they have a tendency to cancel each other rendering the values of $`\mathrm{\Gamma }^{\mathrm{sing}}`$ rather sensitive to the exact values of these two quantities. So far we have only considered OMC for the case of muonic atoms. For the case of a liquid hydrogen target one also has to take into account the possibility of muon capture in a muonic molecule $`p\mu p`$, which can be formed via the reaction $`p\mu +pepp\mu p+e+124`$eV. In such a molecule the muon can be found in a so called ortho $`(O)`$ (spin of the protons parallel) or para $`(P)`$ (spin of the protons antiparallel) spin state relative to its two accompanying protons. The decay rates of these molecular states can be calculated from the singlet/triplet rates of the muonic atom via $`\mathrm{\Gamma }_P`$ $`=`$ $`2\gamma _P{\displaystyle \frac{1}{4}}(3\mathrm{\Gamma }_{\mathrm{trip}}+\mathrm{\Gamma }_{\mathrm{sing}}),`$ (4.47) $`\mathrm{\Gamma }_O`$ $`=`$ $`p_{1/2}\mathrm{\Gamma }_{1/2}+p_{3/2}\mathrm{\Gamma }_{3/2},`$ (4.48) The wavefunction corrections<sup>#14</sup><sup>#14</sup>#14$`\gamma _P(\gamma _O)`$ denotes the ratio of the probability of finding the negative muon at the point occupied by a proton in the para-muonic (ortho-muonic) molecule and the probability of finding the negative muon at the origin in the muonic atom. We are grateful to Shung-Ichi Ando for pointing out ref. to us. are taken to be $`\gamma _O=0.500,\gamma _P=0.5733`$ . We note that the para molecular state is often referred to as the statistical mixture, as it corresponds to the naively expected occupation numbers of the muonic atom. For a precise calculation of the ortho molecular state on the other hand one first has to know the exact probabilities $`p_{1/2},p_{3/2}`$ for the muonic molecule being in a total spin S=(1-1/2)=1/2 or a total spin S=(1+1/2)=3/2 state, with $`p_{1/2}>0.5`$ . The corresponding decay rates are given by $`\mathrm{\Gamma }_{1/2}`$ $`=`$ $`2\gamma _O\left({\displaystyle \frac{3}{4}}\mathrm{\Gamma }_{\mathrm{sing}}+{\displaystyle \frac{1}{4}}\mathrm{\Gamma }_{\mathrm{trip}}\right)`$ (4.49) $`\mathrm{\Gamma }_{3/2}`$ $`=`$ $`2\gamma _O\mathrm{\Gamma }_{\mathrm{trip}}`$ (4.50) Theoretical calculations of the spin-effects in the muonic molecule suggest $`p_{1/2}1,p_{3/2}0`$ which leads to the values: $`\mathrm{\Gamma }_P^{\mathrm{OMC}}`$ $`=`$ $`208\times \mathrm{s}^1`$ (4.51) $`\mathrm{\Gamma }_O^{\mathrm{OMC}}`$ $`=`$ $`493\mathrm{}519\times \mathrm{s}^1.`$ (4.52) where the range given in $`\mathrm{\Gamma }_O^{\mathrm{OMC}}`$ corresponds to $`0.95p_{1/2}1`$. We have allowed here for a 5% uncertainty in the occupation numbers<sup>#15</sup><sup>#15</sup>#15The possibility of very different occupation numbers has been raised by Shung-Ichi Ando during the Chiral Dynamics 2000 conference. This will be reported in a forthcoming paper, see ref.. to show the sensitivity of our results on this quantity. Since $`\mathrm{\Gamma }_{\mathrm{sing}}\mathrm{\Gamma }_{\mathrm{trip}}`$ for OMC, $`\mathrm{\Gamma }_O^{\mathrm{OMC}}`$ turns out to be roughly proportional to $`p_{1/2}`$. We note that our number for capture from the molecular ortho state agrees very well with the most recent measurement $`\mathrm{\Gamma }_O^{\mathrm{exp}.}=(531\pm 33)\times \mathrm{s}^1`$ from Bardin et al. . Let us stress at this point the importance of the new proposed experiment at PSI which will be done with a hydrogen gas target and will thus be independant on these occupation numbers. One will directly measure $`\mathrm{\Gamma }_{sing}`$. With these expressions given above one can now calculate the rate for muon capture in liquid hydrogen for a general scenario, if one knows the relative occupation numbers for the atomic singlet<sup>#16</sup><sup>#16</sup>#16We assume that all muonic atoms initially in the triplet state are effectively converted into the atomic singlet state through collision with hydrogen molecules in the stopping target via the Gershtein-Zeldovich mechanism . $`f_S`$, the molecular ortho $`f_O`$ and the molecular para $`f_P`$ state: $$\mathrm{\Gamma }_{\mathrm{OMC}}^{H_2}=f_S\mathrm{\Gamma }_{\mathrm{sing}}+f_O\mathrm{\Gamma }_O+f_P\mathrm{\Gamma }_P$$ (4.53) For example, in the case of the recent TRIUMF experiment with $`f_S=0.061,f_O=0.854,f_P=0.085`$ one would obtain a total capture rate of $`\mathrm{\Gamma }_{\mathrm{OMC}}^{\mathrm{TRIUMF}}=(504+𝒪(1/M_N^3))\times \mathrm{s}^1`$. ### D Radiative muon capture #### 1 Total capture rates In the static approximation for the W–boson, the pertinent matrix element for RMC decomposes into two terms, $`_{\mu ^{}p\nu _\mu n\gamma }`$ $`=`$ $`\nu _\mu |W_\mu ^+|\mu i{\displaystyle \frac{g^{\mu \nu }}{M_W^2}}\left[n|𝒯Vϵ^{}V_\nu ^{}|pn|𝒯Vϵ^{}A_\nu ^{}|p\right]`$ (4.55) $`+\nu _\mu \gamma |W_\mu ^+|\mu i{\displaystyle \frac{g^{\mu \nu }}{M_W^2}}\left[n|V_\nu ^{}|pn|A_\nu ^{}|p\right],`$ so that its square in the spin–averaged case can be written as a sum of four terms, with both photons coming either from the hadronic or the leptonic side and two mixed terms, i.e. $`{\displaystyle \frac{1}{4}}{\displaystyle \underset{\sigma \sigma ^{}ss^{}\lambda \lambda ^{}}{}}|^{\mathrm{RMC}}|^2`$ $`=`$ $`{\displaystyle \frac{e^2G_F^2V_{ud}^2}{2}}\left[L_{\mu \nu }^{(a)}H_{(d)}^{\mu \nu }+\left({\displaystyle \underset{\lambda \lambda ^{}}{}}L_{\mu \nu }^{(b)}H_{(c)}^{\mu \nu }+L_{\mu \nu }^{(c)}H_{(b)}^{\mu \nu }\right)+L_{\mu \nu }^{(d)}H_{(a)}^{\mu \nu }\right],`$ (4.56) with $`\lambda ,\lambda ^{}`$ the photon helicities. Explicit expressions for the various tensors are not given here because they are lengthy and not illuminating. We also note that standard packages like REDUCE can not be used straightforwardly to obtain these tensors since the cyclicity of the trace in the presence of $`\gamma _5`$ matrices is not fulfilled. The total decay rate is given by: $$\mathrm{\Gamma }_{\mathrm{tot}}=\frac{|\mathrm{\Phi }(0)_{1s}|^2}{16\pi ^4}_0^\pi \mathrm{sin}\theta d\theta _0^{\omega _{\mathrm{max}}}𝑑\omega \omega l_0^{}\left(1\left(\frac{m_\mu \omega (1\mathrm{cos}\theta )}{M_N}\right)\right)\frac{1}{4}\underset{\sigma \sigma ^{}ss^{}\lambda \lambda ^{}}{}|^{\mathrm{RMC}}|^2,$$ (4.57) with $`\omega =k_0`$ the photon energy. The direction of the photon defines the z–direction and $`\theta `$ in Eq.(4.57) is the polar angle of the outgoing lepton with respect to this direction. The maximal photon energy is given by $$\omega _{\mathrm{max}}=m_\mu \left(1+\frac{m_\mu }{2M_N}\right)\left(1+\frac{m_\mu }{M_N}\right)^1.$$ (4.58) Furthermore, the energy of the outgoing lepton follows from energy conservation, $$l_0^{}=m_\mu \omega \frac{m_\mu ^2}{M_N}+\frac{\omega (1\mathrm{cos}\theta )(m_\mu \omega )}{M_N}+𝒪(1/M_N^2).$$ (4.59) First we discuss the (academic) spin-averaged RMC scenario, which allows for a comparison with previous calculations: $`\mathrm{\Gamma }_{\mathrm{spinav}.}^{\mathrm{RMC}}`$ $`=`$ $`\left(66.0+18.7+𝒪(1/M_N^2)\right)\times 10^3s^1=84.7\times 10^3s^1(\mathrm{HBChPT})`$ (4.60) $`\mathrm{\Gamma }_{\mathrm{spinav}.}^{\mathrm{RMC}}`$ $`=`$ $`\left(66.0+20.4+𝒪(1/M_N^2)\right)\times 10^3s^1=86.4\times 10^3s^1(\mathrm{SSE}).`$ (4.61) Both the HBChPT and the SSE results suggest a good convergence for the chiral expansion, as expected from dimensional analysis. Note that the leading order capture rates in both calculations are identical, as $`\mathrm{\Delta }`$(1232) related effects only start at sub-leading order. Our leading order result also agrees<sup>#17</sup><sup>#17</sup>#17Ref. gives $`\mathrm{\Gamma }_{\mathrm{spinav}.}^{\mathrm{RMC},\mathrm{HBChPT}}=61\times 10^3\mathrm{s}^1`$ to leading order. The small difference can be traced back to the different values of some of their input parameters. with the HBChPT calculation of ref.. However, our HBChPT $`𝒪(p^2)`$ correction is nearly 30% larger than the one given in . Comparing with the corresponding $`𝒪(ϵ^2)`$ correction in SSE, we note that $`\mathrm{\Delta }`$(1232) indeed leads to a larger decay RMC decay rate and constitutes a 9% correction to our $`𝒪(p^2)`$ contribution. However, the total (spin-averaged) decay rate is only affected by 2% due to the fast convergence of the chiral series for RMC. For the case of muonic atoms we obtain the following decay rates in the singlet/triplet channel<sup>#18</sup><sup>#18</sup>#18Note the reversal of the relative size of the singlet to triplet contribution as compared to the case of OMC., utilizing the projection formalism outlined in sec.IV C. $`\mathrm{\Gamma }_{\mathrm{sing}}^{\mathrm{RMC}}`$ $`=`$ $`\left(12.718.7\mathrm{GeV}/M_N+𝒪(1/M_N^2)\right)\times 10^3s^1=3.10\times 10^3s^1(\mathrm{HBChPT})`$ (4.62) $`\mathrm{\Gamma }_{\mathrm{sing}}^{\mathrm{RMC}}`$ $`=`$ $`\left(12.718.3\mathrm{GeV}/M_N+𝒪(1/M_N^2)\right)\times 10^3s^1=2.90\times 10^3s^1(\mathrm{SSE})`$ (4.63) $`\mathrm{\Gamma }_{\mathrm{trip}}^{\mathrm{RMC}}`$ $`=`$ $`\left(1193.86\mathrm{GeV}/M_N+𝒪(1/M_N^2)\right)\times 10^3s^1=112\times 10^3s^1(\mathrm{HBChPT})`$ (4.65) $`\mathrm{\Gamma }_{\mathrm{trip}}^{\mathrm{RMC}}`$ $`=`$ $`\left(1191.80\mathrm{GeV}/M_N+𝒪(1/M_N^2)\right)\times 10^3s^1=114\times 10^3s^1(\mathrm{SSE})`$ (4.66) Note that for the total numbers given we did not expand all kinematical factors in powers of $`1/M_N`$ since in case of the small singlet, the contribution from the terms starting at order $`1/M_N^2`$ can not be neglected. In fact, a strict truncation at $`1/M_N`$ leads to an unphysical negative singlet capture rate. For the much bigger triplet, these higher order corrections are much less important. We remark that in ref. no strict $`1/M_N`$ expansion was performed, only at some places these authors used the leading order results, e.g. for the nucleon energy by neglecting the recoil term. Only if one goes to a sufficiently high order in the small scale expansion, the truncation of these kinematical factors can be justified. Comparing the $`𝒪(p^2)`$ HBChPT with the $`𝒪(ϵ^2)`$ SSE calculation, we observe that while the total singlet capture rate is nearly identical in both approaches, the absolute value of the $`1/M_N`$ term in the total triplet capture rate is a factor of two different between SSE and HBChPT. However this NLO term is much smaller than the LO triplet rate, leading to a rather similar total spin triplet RMC rate with or without explicit $`\mathrm{\Delta }`$(1232) contributions. Finally we address the complications for RMC due to the presence of muonic molecules in the liquid hydrogen target. According to Eq.(4.47), we can easily determine the capture rate from the molecular para state $`\mathrm{\Gamma }_P^{\mathrm{RMC}}`$ $`=`$ $`85.2\times 10^3s^1(\mathrm{HBChPT})`$ (4.67) $`\mathrm{\Gamma }_P^{\mathrm{RMC}}`$ $`=`$ $`86.4\times 10^3s^1(\mathrm{SSE}).`$ (4.68) Let us now turn to the molecular ortho state, which turns out to dominate in the recent RMC experiment from TRIUMF . One obtains for $`p_{1/2}=1`$: $`\mathrm{\Gamma }_O^{\mathrm{RMC}}`$ $`=`$ $`30.4\times 10^3s^1(\mathrm{HBChPT})`$ (4.69) $`\mathrm{\Gamma }_O^{\mathrm{RMC}}`$ $`=`$ $`30.8\times 10^3s^1(\mathrm{SSE}).`$ (4.70) Due to the triplet dominance in RMC (as opposed to the singlet dominance in OMC) $`\mathrm{\Gamma }_O^{\mathrm{RMC}}`$ is now roughly proportional to $`(13/4p_{1/2})`$ which leads to a big sensitivity of the RMC capture rate to the exact occupation numbers of the relative molecular sub-states. For example, a 5% uncertainty in the occupation numbers $`p_{1/2}=0.95,p_{3/2}=0.05`$ would lead to a 13% change in the ortho capture rate $`\mathrm{\Gamma }_O^{\mathrm{RMC}}35\times 10^3s^1`$. We will discuss the implications of this uncertainty when we compare our results with the measured photon spectrum from TRIUMF in the next section. For the total capture rate in the TRIUMF experiment $$\mathrm{\Gamma }_{\mathrm{RMC}}^{H_2}=f_S\mathrm{\Gamma }_{\mathrm{sing}}^{\mathrm{RMC}}+f_O\mathrm{\Gamma }_O^{\mathrm{RMC}}+f_P\mathrm{\Gamma }_P^{\mathrm{RMC}}$$ (4.71) with $`f_S=0.061,f_O=0.854,f_P=0.085`$ one would obtain $`\mathrm{\Gamma }_{\mathrm{RMC}}^{TRIUMF}=(34.3[34.8]+𝒪(1/M_N^2))\times 10^3\mathrm{s}^1`$ in HBChPT \[SSE\]. This leads to a relative branching ratio $`Q_\gamma =\mathrm{\Gamma }_{\mathrm{RMC}}/\mathrm{\Gamma }_{\mathrm{OMC}}`$ $`Q_\gamma ^{\mathrm{HBChPT}}`$ $`=`$ $`{\displaystyle \frac{34.3\times 10^3\mathrm{s}^1}{504\mathrm{s}^1}}=6.8\times 10^5+𝒪(1/M_N^2))`$ (4.72) $`Q_\gamma ^{\mathrm{SSE}}`$ $`=`$ $`{\displaystyle \frac{34.8\times 10^3\mathrm{s}^1}{504\mathrm{s}^1}}=6.9\times 10^5+𝒪(1/M_N^2)).`$ (4.73) Unfortunately the full relative branching ratio is not accessible in experiment, as one has to use a severe cut on the photon energies due to strong backgrounds. In the TRIUMF experiment only photons with an energy $`\omega >60`$MeV were detected. We therefore now move on to a discussion on the photon spectrum. #### 2 Photon spectrum The photon spectrum $`d\mathrm{\Gamma }/d\omega `$ can be obtained straightforwardly from Eq.(4.57). We refrain from giving the lengthy formulae for the various atomic states here. We have calculated the photon spectra to $`𝒪(p^2)`$ in HBChPT and to $`𝒪(ϵ^2)`$ in SSE. The resulting curves are very similar. In fig.4 we show the SSE $`𝒪(ϵ^2)`$ results with the coupling values $`g_{\pi N\mathrm{\Delta }}\times b_1=1.05\times 12=12.6`$ for the singlet, triplet, para and ortho states. The relative difference between the $`𝒪(p^2)`$ HBChPT and the $`𝒪(ϵ^2)`$ SSE calculation for all states are shown in fig.5, showing explicitly the small role of $`\mathrm{\Delta }`$(1232) in RMC. With the exception of the small singlet, the $`\mathrm{\Delta }`$ effects amount to less than 5% for all photon energies. Only in the case of the singlet, a more pronounced influence of the $`\mathrm{\Delta }`$ is observed. We note that these results are very similar to the ones found by Beder and Fearing for the spectra and the relative contribution from the spin–3/2 resonance, although their calculation is based on a very different approach. Even the result for the singlet is comparable though not identical to the one of Beder and Fearing. It changes sign for photon energies of about 74 MeV. Like for the case of the small triplet in OMC, it is expected that the small singlet (for RMC) is more sensitive to $`1/M_N`$ corrections. We have also increased the coupling $`b_1`$ to values of 24 and 60 thus enhancing the $`\mathrm{\Delta }`$(1232) contributions in the SSE calculation by a factor of 2 and 5, respectively. We find very little sensitivity to this, e.g. the maximum in the photon spectrum for the para state changes from 1.47 GeV$`{}_{}{}^{1}\mathrm{s}_{}^{1}`$ for $`b_1=12`$ to 1.55 GeV$`{}_{}{}^{1}\mathrm{s}_{}^{1}`$ for $`b_1=60`$. Correspondingly, the total decay rate for the para state is increased by 7%. The smallness of the $`\mathrm{\Delta }`$(1232) related effects has two reasons. First, as noted already, it only appears at NLO, whereas the RMC process is mainly controled by the leading order effects, hinting to a good convergence behavior of the chiral expansions. Second, its sole contribution at that order is in axial–vector correlator, but not in the other three correlators. This already leads to a statistical suppression as compared to the pure nucleonic contributions. Third, despite the largeness of the coupling $`g_{\pi N\mathrm{\Delta }}\times b_1`$ (remember that $`b_1`$ is related to the dominant $`M1`$ $`N\mathrm{\Delta }`$ transition ), the delta contribution to the VA–correlator is still smaller than the one from the nucleon, which is enhanced by the large isovector magnetic moment. The dominant axial and axial–vector contribution comes indeed from the leading pion pole, which is not affected by the $`\mathrm{\Delta }`$(1232) effects. To quantify these statements, let us for a moment consider the $`𝒪(p^2)`$ HBChPT calculation. If one switches off the complete contribution from the axial correlator, Eq.(4.4), the singlet rate is enhanced by a factor of 1.7, whereas the triplet is decreased by a factor of about 170 ! Consequently, we then have $`\mathrm{\Gamma }_{\mathrm{tot}}7\times 10^3\mathrm{s}^1`$, which is an order of magnitude smaller than the value given in the previous section. If one on the other hand sets $`\kappa _s=\kappa _v=0`$ in the V, VV and AV correlators, the singlet rate increases by a factor of about 2.7 and the triplet rate decreases by a factor of 1.3, leading to a total rate of 68$`\times 10^3\mathrm{s}^1`$. It is also instructive to consider the chiral limit, $`m_\pi =0`$. To be specific, we discuss the $`𝒪(p^2)`$ HBChPT calculation, with the $`𝒪(ϵ^2)`$ SSE results being very similar. To leading order in $`1/M_N`$, one encounters a pole at $`\omega =m_\mu /2`$ in the A and AV correlators. This can be seen by looking at pion pole terms in the chiral limit, which take the form $$\frac{1}{(ll^{})^2}=\frac{1}{m_\mu ^22m_\mu l_0^{}},$$ (4.74) and using the leading order result $`l_0^{}=m_\mu \omega +𝒪(1/M_N)`$, cf. Eq.(4.59). This well-known Bethe–Heitler pole is independent of the angular variable $`x`$. The condition for this pole to appear is that the pion mass has to be below the muon mass. For zero pion mass, there is another pole at the same energy for $`x=1`$ stemming from the terms $`(rr^{})^1`$. Furthermore, the chiral limit photon spectra of the $`𝒪(p^2)`$ HBChPT calculation are shown in Fig.6. We remark that these spectra are very different from the ones with the physical pion mass due to the abovementioned singularities. We also point out that the pion mass effects are larger in RMC than in the OMC case discussed above. In particular, setting e.g. $`m_\pi =125`$MeV, the rate in the para state (total rate) increases to 99$`\times 10^3\mathrm{s}^1`$. We will discuss the implications on the TRIUMF measurement in the following section. #### 3 Discussion of the TRIUMF result for $`g_P`$ The photon spectra discussed in section IV D 2 allow in principle to determine the induced pseudoscalar form factor. The TRIUMF result for $`g_P`$ is obtained by multiplying the terms proportional to the pseudoscalar form factor with a constant denoted $`R`$ (the momentum dependence assumed to be entirely given by the pion pole). The value of $`R`$ is then extracted using the model of Fearing et al. to match the partial rate for photon energies larger than 60 MeV. If we perform such a procedure, we get a similar shift in the partial photon spectra (using the same weight factors for the various $`\mu p`$ states as given in ref.). It is, however, obvious from our analysis that such a procedure is not legitimate. By artificially enhancing the contribution $`g_P`$ (to simulate this procedure, we have introduced the factor $`R`$ in Eqs.(4.4,4.23)), one mocks up a whole class of new contact and other terms not present in the Born term model. To demonstrate these points in a more quantitative fashion, we show in fig.8 the partial branching fraction for our calculation in comparison to the one with $`g_P`$ enhanced by a factor 1.5 and a third curve, which is obtained by increasing $`g_P`$ only by 15%, take $`b_1=24`$ (i.e. enhancing this coupling by a factor of two) and use $`\mathrm{\Delta }=273`$MeV, since in the dispersion theoretical analysis of pion–nucleon scattering the pole in the $`P_{33}`$ partial wave is located at $`W=1210`$MeV. This is shown in fig.8 by the dashed line and it shows that such a combination of small effects can explain most (but not all) of the shift in the spectrum. This is further sharpened by using now the neutral pion mass of 134.97 MeV instead of the charged pion mass, leading to the dotted curve in fig.8. Since the pion mass difference is almost entirely of electromagnetic origin, one might speculate that isospin–breaking effects should not be neglected (as done here and all other existing calculations). Furthermore, as discussed above a slight change in the occupation numbers $`p_{1/2}`$ and $`p_{3/2}`$ would also lead to an increase in the ortho capture rate which could close the gap between the empirical and theoretical results. For example the dashed curve in fig.8 would be moved from 0.42 to 0.48 while the dotted one would go from 0.50 to 0.54 with $`p_{1/2}=0.95`$ and $`p_{3/2}=0.05`$. The situation is reminiscent of the sigma term analysis, where many small effects combine to give the sizeable difference between the sigma term at zero momentum transfer and at the Cheng-Dashen point. We further point out that although the present investigation combined with the findings in ref. does not seem to give any large new term at order $`ϵ^2`$ or from the chiral loops at third order, it can not be excluded that one–loop graphs with insertions from the dimension two chiral Lagrangian (which are formally of fourth order) can give rise to larger corrections than the third order loop and tree terms calculated in ref.. In fact, as we noted before, the photon spectrum is more sensitive to changes in the anomalous magnetic moments than to the induced pseudoscalar coupling. Therefore it can be speculated that one–loop graphs with exactly one insertion $`\kappa _v`$ can generate large corrections. This appears plausible but needs to be supported by a fourth order calculation. Such a mechanism would, however, be much more natural than the simple rescaling of $`g_P`$ based on tree level diagrams only. Another point against this simple rescaling comes from OMC. Indeed if this rescaling holds for RMC it should also hold for OMC. We thus have performed a similar calculation in OMC. Taking the same value for R, one would obtain $`\mathrm{\Gamma }_{\mathrm{OMC}}^{R=1.5}=172.8\times s^1,\mathrm{\Gamma }_{\mathrm{OMC}}^{\mathrm{sing},\mathrm{R}=1.5}=634.6\times s^1,\mathrm{\Gamma }_{\mathrm{OMC}}^{\mathrm{trip},\mathrm{R}=1.5}=18.9\times s^1,`$ (4.75) leading to $`\mathrm{\Gamma }_O^{\mathrm{OMC},\mathrm{R}=1.5}=477\times s^1`$, which is lower than the error bars of the experimental result from Bardin et al. . As expected, the singlet and triplet capture rates are much more sensitive to the details of the interaction than the total rate. As a conclusion we note that the effect of enhancing the capture rates in RMC via setting $`R=1.5`$ leads to a strong reduction of the corresponding OMC rates leading to conflicts with the experimentally determined ortho capture rate. To summarize this discsussion, we have pointed out that two effects in particular have to be investigated in more detail: * the occupation numbers of the atomic structure in muonic atoms/molecules need to be carefully re-examined by experts in this field. * the N<sup>2</sup>LO calculation should be redone including all isospin breaking effects because of the sensitivity to the exact pion mass in the pion-pole contributions, for example. The sum of these small effects should explain the observed photon spectrum, as we believe that the proper hadronic/weak physics part is well under control by now, as our analysis has re-confirmed. A simple rescaling of the pseudoscalar coupling constant should no longer be considered. ## V Summary In this manuscript, we have considered ordinary and radiative muon capture on the proton in the framework of the small scale expansion to third and second order in small momenta, $`𝒪(ϵ^3)`$ and $`𝒪(ϵ^2)`$, respectively. We have also discussed the induced pseudoscalar form factor of the nucleon and its determination from the TRIUMF RMC data. The pertinent results of this investigation can be summarized as follows: * To third order in the small scale expansion, ordinary muon capture is almost not affected by $`\mathrm{\Delta }`$(1232) isobar effects. The only effect comes via the Pauli radius and is extremely small. The NLO contribution to the total capture rate amounts to a 25% correction of the leading term. This is in agreement with naive dimensional counting, which lets one expect corrections of the size $`m_\mu /\mathrm{\Lambda }_\chi `$. This calculation involves very few and well-controlled parameters. We argued that the formula derived from a relativistic calculation, see Eq.(4.45), and extensively used in the literature does not hold in the modern view point of power counting. We have stressed the importance of the upcoming PSI experiment. * To second order in the small scale expansion, we have considered radiative muon capture. $`\mathrm{\Delta }`$(1232) related effects only appear at NLO and its effects on the total capture rate and the photon spectrum are of the order of a few percent. The smallness of the $`\mathrm{\Delta }`$(1232) contributions is due to a combination of effects as discussed in section IV D 2. This agrees with earlier findings in a more phenomenological approach . Isobar effects can therefore not resolve the discrepancy between the TRIUMF measurement for the partial decay width $`\mathrm{\Gamma }(\omega >60`$MeV) and the theoretical predictions. We have, however, pointed out severe loopholes concerning the extraction of $`g_P`$ as done in ref., one of them being the contradiction with the OMC data. In our opinion the most probable explanation of the discrepancy is a combination of many small effects, as detailed in section IV D 3. * The induced pseudoscalar form factor measured in charged pion electroproduction is not very well determined, but clearly is in agreement with the one–loop chiral perturbation theory prediction . A more precise measurement for small invariant momentum transfer squared is called for . ## VI Acknowledgments We would like to thank V. Markushin for useful discussions. We are grateful to Shung-Ichi Ando, Fred Myhrer and Kuniharu Kubodera for communicating the results of ref. prior to publication. ## A Feynman rules In this appendix, we collect the pertinent Feynman rules for calculating OMC and RMC. These read: Isovector vector source in ($`q_\mu `$), nucleon: $$iv𝐯+\frac{i}{2M_0}\left((r_i+r_f)𝐯(r_i+r_f)vv𝐯\right)+\frac{i}{M_0}(1+\dot{\kappa }_v)[S𝐯,Sq],$$ (A.1) Isovector axial source in, nucleon: $$i\mathrm{\hspace{0.17em}2}\dot{g}_AS𝐚i\frac{\dot{g}_A}{M_0}S(r_i+r_f)v𝐚,$$ (A.2) Isovector axial source in, pion out ($`k_\mu `$): $$F_0\frac{1}{2}\mathrm{Tr}[\tau ^i𝐚^{}k+𝐚k\tau ^i],$$ (A.3) Isovector axial source, vector source, pion: $$\frac{F_0}{2}\mathrm{Tr}\{[𝐯_\mu ,\tau ^i]𝐚^\mu +𝐚_\mu [𝐯^\mu ,\tau ^i]\},$$ (A.4) 2 vector sources in, nucleon: $$\frac{i}{2M_0}\left\{(𝐯+v^{(s)})^2\left[v(𝐯+v^{(s)})\right]^2\left(1+\dot{\kappa }_v\right)[S^\mu ,S^\nu ][𝐯_\mu ,𝐯_\nu ]\right\},$$ (A.5) vector source and isovector axial source in, nucleon: $$i\frac{\dot{g}_A}{M_0}\left[S\left(𝐯+v^{(s)}\right)v𝐚+v𝐚S\left(𝐯+v^{(s)}\right)\right],$$ (A.6) $`\mathrm{\Delta }_\mu ^i`$ in, nucleon and vector source ($`k_\mu `$) out: $$+i\frac{b_1}{2M_0}\mathrm{Tr}[\tau ^i(k^\mu 𝐯^\nu k^\nu 𝐯^\mu )]S_\nu ,$$ (A.7) Nucleon in, $`\mathrm{\Delta }_\mu ^i`$ and vector source ($`k_\mu `$) out: $$i\frac{b_1}{2M_0}\mathrm{Tr}[\tau ^i(k^\mu 𝐯^\nu k^\nu 𝐯^\mu )]S_\nu ,$$ (A.8) Isovector axial source, nucleon, $`\mathrm{\Delta }_\mu ^i`$: $$ig_{\pi N\mathrm{\Delta }}\mathrm{Tr}[\tau ^i𝐚_\mu ].$$ (A.9) Tables Figures
warning/0001/math0001039.html
ar5iv
text
# Polynomial hulls and 𝐻^∞ control for a hypoconvex constraint
warning/0001/cond-mat0001405.html
ar5iv
text
# HD-THEP-00-02WUE-ITP-2000-007 Learning structured data from unspecific reinforcement ## 1 Introduction Learning from unspecific reinforcement may be essential in various contexts, both natural and artificial, where, typically, the results of particular actions add to a final consequence which only is valuated. The freedom residing in each step is not (or only partially) controlled directly and the learner must cope with the necessity of improving its performance only from information concerning the final success of a complex series of actions. It is therefore important to find out whether there are simple and robust procedures for such situations, which might have developed under natural conditions and which may be basic also for artificial learning rules (for this reason we do not consider evolved algorithms like Q-learning , TD learning etc). In previous works , we have introduced an “Association-Reinforcement” learning model based on the following conception: 1. For each given input (external situation) the agent answers with an action (operation) depending solely on the input and on its instantaneous internal (cognitive) structure and simultaneously strengthens (in its internal structure) the blind association between this particular input and action. 2. At the end of a series of actions (path) the final success is judged. Then the associations “situation - operation” which have been involved on this path are re-weighted equally and depending only on the final success – unspecific reinforcement. In we studied an implementation of this model to a classification problem for perceptrons, the Association-Reinforcement-Hebb-rule and showed some amazing properties: a) Despite the fact that feedback on the learner’s performance enters its learning dynamics only in an unspecific way in that it cannot be associated with single identifiable correct or incorrect associations, convergence of the AR-Hebb-algorithm in the sense of asymptotically perfect generalization is found. b) For given initial conditions, this convergence depends on the learning parameters characterizing the 2 steps described above; in particular none of these steps can be completely inhibited. Alternatively, for given algorithm parameters convergence may depend on initial conditions. In detail the dynamics of this algorithm was found to be very complex and interesting, being controlled by fixed points in the pre-asymptotic regime, and having a continuous set of asymptotic convergence laws. These results could easily be extended to the more realistic case where in the second step the unspecific reinforcement is randomly applied to only part of the associations achieved in the first step (the agent does not recall everything it has done on the trial) . Further interesting extensions concern the question of structured data and of multi-layer perceptrons. Structured data represent a more involved classification problem and it is known that when teacher and data vector are not fully aligned (or exactly uncorrelated) the usual Hebb rule does not lead to convergence of the student vector onto that of the teacher, while the perceptron algorithm does . On the other hand, the limiting case of the AR-Hebb-rule corresponding to the perceptron rule has been shown not to converge in the case of unspecific reinforcement for non-structured data. It is therefore a non-trivial question whether the unspecific reinforcement problem can be solved for structured data and in particular, whether some immediate extension of the AR-Hebb-rule can be shown to converge in this case. It is this question which we shall address in this paper. In a future publication we shall treat the problem of the committee machine as a first step to multi-layer perceptrons. In section 2 we shall describe the learning model in the general setting and in section 3 we shall discuss its convergence properties providing numerical and analytical results. Thereby we shall briefly recall the non-structured data case and then concentrate on the general, structured data case. Section 4 contains the conclusions. ## 2 Learning rule for perceptrons under unspecific reinforcement We consider one layer perceptrons with Ising or real number units $`s_i`$, real weights (synapses) $`J_i`$ and one Ising output unit: $$s=\mathrm{sign}\left(\frac{1}{\sqrt{N}}\underset{i=1}{\overset{N}{}}J_is_i\right)$$ (1) Here $`N`$ is the number of input nodes, and we put no explicit thresholds. The network (student) is presented with a series of patterns $`s_i=\xi _i^{(q,l)}`$, $`q=1,\mathrm{},Q`$, $`l=1,\mathrm{},L`$ to which it answers with $`s^{(q,l)}`$. A training period consists of the successive presentation of $`L`$ patterns. The answers are compared with the corresponding answers $`t^{(q,l)}`$ of a teacher with pre-given weights $`B_i`$ and the average error made by the student over one training period is calculated: $$e_q=\frac{1}{2L}\underset{l=1}{\overset{L}{}}|t^{(q,l)}s^{(q,l)}|.$$ (2) The training algorithm consists of two parts: * \- a “blind” Hebb-type association at each presentation of a pattern: $$J_{i}^{}{}_{}{}^{(q,l+1)}=J_{i}^{}{}_{}{}^{(q,l)}+\frac{a_1}{\sqrt{N}}s^{(q,l)}\xi _i^{(q,l)};$$ (3) * \- an “unspecific” but graded reinforcement proportional to the average error $`e_q`$ introduced in (2), also Hebbian, at the end of each training period, $$J_{i}^{}{}_{}{}^{(q+1,1)}=J_i^{(q,L+1)}\frac{a_2}{\sqrt{N}}e_q\underset{l=1}{\overset{L}{}}r_ls^{(q,l)}\xi _i^{(q,l)}.$$ (4) where $`e_q`$ is the average error eq. (2) and $`r_l`$ is a dichotomic random variable: $$r_l=\{\begin{array}{ccc}1\hfill & \mathrm{with}\mathrm{probability}\hfill & w\hfill \\ 0\hfill & \mathrm{with}\mathrm{probability}\hfill & 1w\hfill \end{array}$$ (5) Because of these 2 steps we called this algorithm “association/reinforcement(AR)-Hebb-rule”. We are interested in the behavior with the number of iterations $`q`$ of the generalization error $`ϵ_g(q)`$: $$ϵ_g(q)=\frac{1}{\pi }\mathrm{arccos}\left(\frac{JB}{|J||B|}\right),$$ (6) in particular we shall test whether the behavior of $`ϵ_g(q)`$ follows a power law at large $`q`$: $$ϵ_g(q)\mathrm{const}q^p.$$ (7) The training patterns $`\{\xi _i^{(q,l)}\}`$ are generated randomly from the following distribution: $`P(\xi )`$ $`=`$ $`{\displaystyle \frac{1}{2}}{\displaystyle \underset{\sigma =\pm 1}{}}P(\xi |\sigma )`$ $`P(\xi |\sigma )`$ $`=`$ $`{\displaystyle \underset{i=1}{\overset{N}{}}}{\displaystyle \frac{1}{\sqrt{2\pi }}}\mathrm{e}^{\frac{1}{2}(\xi _im\sigma C_i)^2}`$ (8) and we take: $$C^2=B^2=N,CB=\eta N$$ (9) with fixed, given $`m,\eta `$. Notice the following features: * During training the student only uses its own associations $`\xi ^{(q,l)}s^{(q,l)}`$ and the average error $`e_q`$ which does not refer specifically to the particular steps $`l`$. * Since the answers $`s^{(q,l)}`$ are made on the basis of the instantaneous weight values $`J^{(q,l)}`$ which change at each step according to eq. (3), the series of answers form a correlated sequence with each step depending on the previous one. Therefore $`e_q`$ measures in fact the performance of a “path”, an interdependent set of decisions. * In contrast with the case studied in the patterns can now have a structure. This introduces essential differences to the previous situation, as we shall see in the next section. * We explicitly account for imperfect recall at the reinforcement step by the parameter $`w`$ (5). This introduces a supplementary, biologically motivated randomness which, as already suggested in , does not appear to introduce qualitative changes in the results, however (see section 3). * For $`L=1`$ (and $`w=1`$) the algorithm reproduces the usual “perceptron rule” (for $`a_1=0`$) or to the usual “unsupervised Hebb rule” (for $`a_2=2a_1`$) for on-line learning, for which the corresponding asymptotic behavior is known , , . To study the learning behaviour we use Monte Carlo simulation and coarse grained analysis. The latter is provided by combining the blind association (3) during a learning period of $`L`$ elementary steps and the graded unspecific reinforcement (4) at the end of each learning period into one coarse grained step $`J_i^{(q+1,1)}=J_i^{(q,1)}+{\displaystyle \frac{1}{\sqrt{N}}}(a_1a_2e_q){\displaystyle \underset{l=1}{\overset{L}{}}}r_l\mathrm{sign}\left({\displaystyle \frac{1}{\sqrt{N}}}{\displaystyle \underset{j=1}{\overset{N}{}}}J_j\xi _j^{(q,l)}\right)\xi _i^{(q,l)},`$ (10) $`e_q={\displaystyle \frac{1}{2L}}{\displaystyle \underset{l=1}{\overset{L}{}}}\left|\mathrm{sign}\left({\displaystyle \frac{1}{\sqrt{N}}}{\displaystyle \underset{k=1}{\overset{N}{}}}J_k\xi _k^{(q,l)}\right)\mathrm{sign}\left({\displaystyle \frac{1}{\sqrt{N}}}{\displaystyle \underset{i=k}{\overset{N}{}}}B_k\xi _k^{(q,l)}\right)\right|.`$ (11) For simplicity we shall take for the time being $`r_l=1`$, i. e. $`w=1`$ in eq. (5). We introduce $$\alpha =qL/N,\lambda =a_1/a_2$$ (12) and rescale everything with $`a_2`$, which means that we can take without loss of generality $`a_2=1`$ in (10),(11). We define the overlaps: $$(\alpha )=\frac{1}{N}BJ^{(q,l)},𝒬(\alpha )=\frac{1}{N}[J^{(q,l)}]^2,𝒟(\alpha )=\frac{1}{N}CJ^{(q,l)}.$$ (13) Note that in the “thermodynamic limit” $`L/N0`$ the overlaps are self-averaging and we can neglect the dependence of $``$, $`𝒟`$ and $`𝒬`$ on $`l`$. We shall follow standard procedures , , , . Treating $`\alpha `$ as a continuous variable and using: $`x`$ $``$ $`\pi ϵ_g=\mathrm{arccos}\left({\displaystyle \frac{}{\sqrt{𝒬}}}\right),`$ (14) $`y`$ $``$ $`\mathrm{arccos}\left({\displaystyle \frac{𝒟}{\sqrt{𝒬}}}\right),`$ (15) $`z`$ $``$ $`\mathrm{arccos}\eta `$ (16) we obtain the coarse grained equations: $`{\displaystyle \frac{d}{d\alpha }}`$ $`=`$ $`{\displaystyle \frac{d}{d\alpha }}\left(\sqrt{𝒬}\mathrm{cos}x\right)=\left(\lambda {\displaystyle \frac{1}{2}}\right)A_{JT}+{\displaystyle \frac{1}{2L}}A_{TT}+{\displaystyle \frac{1}{2}}\left(1{\displaystyle \frac{1}{L}}\right)S_{JT}A_{JT},`$ (17) $`{\displaystyle \frac{d𝒟}{d\alpha }}`$ $`=`$ $`{\displaystyle \frac{d}{d\alpha }}\left(\sqrt{𝒬}\mathrm{cos}y\right)=\left(\lambda {\displaystyle \frac{1}{2}}\right)A_{JC}+{\displaystyle \frac{1}{2L}}A_{TC}+{\displaystyle \frac{1}{2}}\left(1{\displaystyle \frac{1}{L}}\right)S_{JT}A_{JC},`$ (18) $`{\displaystyle \frac{d\sqrt{𝒬}}{d\alpha }}`$ $`=`$ $`\left(\lambda {\displaystyle \frac{1}{2}}\right)A_{JJ}+{\displaystyle \frac{1}{2L}}A_{TJ}+{\displaystyle \frac{1}{2}}\left(1{\displaystyle \frac{1}{L}}\right)S_{JT}A_{JJ}+`$ (19) $`{\displaystyle \frac{1}{2\sqrt{𝒬}}}\left[\left(\lambda {\displaystyle \frac{1}{2}}\right)S_{JT}+{\displaystyle \frac{1}{4}}\left(1{\displaystyle \frac{1}{L}}\right)S_{JT}^2+\left(\lambda {\displaystyle \frac{1}{2}}\right)^2+{\displaystyle \frac{1}{4L}}\right],`$ where the expectation values $`A_.,S_.`$ are given in Appendix (section 5.1). These equations describe the flow of the three quantities $`ϵ_g=x/\pi `$, $`𝒬`$ and $`y`$ with $`\alpha `$ and involve the data/teacher parameters $`m`$ and $`z=\mathrm{arcos}\eta `$ and the learning parameter $`\lambda `$. Note the geometric constraint: $$\mathrm{sin}\frac{y}{2}\mathrm{sin}\frac{z}{2}=\omega \mathrm{sin}\frac{x}{2},|\omega |1.$$ (20) ## 3 Convergence behaviour of the AR-Hebb algorithm ### 3.1 Non-structured data The case of non-structured data – $`m=0`$ in eq. (35)-(43) – has been treated in , here we briefly recall some of the results for the later comparison with the structured data case. Monte Carlo simulations indicate that in spite of the partial information contained in the unspecific reinforcement perfect generalization is achieved by the AR-Hebb algorithm and it depends on the learning parameters – see . This intriguing behaviour is elucidated by the coarse grained analysis. In this case eqs. (17)-(19) reduce to two equations (for $``$ and $`𝒬`$) which have as general asymptotic solutions $`ϵ_g^2`$ $``$ $`{\displaystyle \frac{1}{2\pi (\frac{1}{\lambda L}1)}}\alpha ^1+\stackrel{~}{c}_1\alpha ^{\frac{1}{\lambda L}}\mathrm{for}\lambda {\displaystyle \frac{1}{L}},`$ (21) $`ϵ_g^2`$ $``$ $`\left({\displaystyle \frac{1}{2\pi }}\mathrm{ln}\alpha +\stackrel{~}{c}_2\right)\alpha ^1\mathrm{for}\lambda ={\displaystyle \frac{1}{L}},`$ (22) $`𝒬`$ $``$ $`{\displaystyle \frac{2}{\pi }}\lambda ^2\alpha ^2`$ (23) at large $`\alpha `$. We see that for $`\lambda <\frac{1}{L}`$ we obtain asymptotically perfect generalization, the dominant term exhibiting the usual power -1/2 , while for $`\lambda >\frac{1}{L}`$ the second term in (21) dominates and ensures again perfect generalization but with a different power law, $`1/(2\lambda L)`$. For $`\lambda =\frac{1}{L}`$ we obtain logarithmic corrections – see eq. (22). Notice that these results hold also for $`L=1`$. One can generally see that for $`\lambda =0`$ one cannot have perfect generalization for $`L>1`$. For $`L=1`$ one re-obtains the asymptotic behavior found in . This learning algorithm is further characterized by highly interesting pre-asymptotics, dominated by two stationarity conditions, one for the self-overlap, $`d𝒬/d\alpha =0`$, and one for the generalization error $`dϵ_g/d\alpha =0`$. For suitable values of the network parameters, the two stationarity conditions may simultaneously be satisfied, leading to fixed points of the learning dynamics, one of them stable and of poor generalization, the other with one attractive and one repulsive direction. Correspondingly, the flow is divided by a separatrix defined by a critical $`\lambda _c(𝒬_0)`$ into trajectories leading to convergence according to the asymptotic behaviour (21)-(23) for $`\lambda >\lambda _c(𝒬_0)`$, or to poor generalization otherwise. The salient features of these results for the case of non-structured data are the convergence of the AR-Hebb-algorithm in the sense of asymptotically perfect generalization with a power law depending on the learning parameters $`L`$ and $`\lambda `$ and the existence of a minimal value $`\lambda _c(𝒬_0)`$, fixed by the pre-asymptotic structure and below which the system is driven toward complete confusion. Notice also that the best convergence is achieved for $`\lambda `$ just above $`\lambda _c`$. One last point concerns the recalling parameter $`w`$, eqs. (4),(5). A rough first quantitative characterization of this modification would be that it leads to an effective rescaling of the parameter $`\lambda `$, viz. $`\lambda \lambda /w`$, leading to a corresponding reduction of critical $`\lambda `$’s by approximately a factor $`p`$. This is well supported by numerical simulations (see also Fig. 1 for the case of structured data) and we conclude that the algorithm is stable against this supplementary element of indeterminism. ### 3.2 Structured data Numerical simulations indicate that for $`m0`$ and $`0<|\eta |<1`$ the behaviour of the algorithm for all $`w`$ is more involved: generically, no convergence is found in this case for fixed values of the learning parameters. This agrees with the expectations, since, on the one hand the situation found at $`L=1`$, $`w=1`$ for structured data could be expected to hold for every $`L`$, namely that Hebb updating leads to a nonzero asymptotic generalization error. On the other hand, the situation found before for non-structured data should hold also for structured data, namely that the perceptron rule ($`\lambda =0`$) (which for $`L=1`$ was shown to lead to convergence also in the structured data case ) does not work for $`L>1`$. In fact one can make a more general argument that for fixed $`\lambda `$ the AR-Hebb rule does not lead to perfect generalization for generically structured data. To obtain good generalization requires $`/\sqrt{𝒬}1`$ and $`𝒟/\sqrt{𝒬}\eta `$, from which one may obtain the necessary dominant scaling (with $`𝒬`$) of the various integrals appearing in (35)-(43), namely $`A_{JT}`$ $``$ $`\kappa ,`$ $`A_{TJ}`$ $``$ $`\sqrt{Q}\kappa +o(\sqrt{Q}),`$ $`A_{JJ}`$ $``$ $`\sqrt{Q}\kappa +o(\sqrt{Q}),`$ $`A_{JC}`$ $``$ $`\kappa ,`$ $`A_{TC}`$ $``$ $`\kappa ,`$ $`A_{TT}`$ $``$ $`\kappa ,`$ $`S_{JT}`$ $``$ $`1,`$ with $$\kappa =m\eta \phi (m\eta )+\sqrt{\frac{2}{\pi }}\mathrm{e}^{m^2\eta ^2/2}.$$ (24) This in turn would lead to the following asymptotic expressions for the flow equations (17)-(19) (at fixed $`\lambda `$) $`{\displaystyle \frac{d}{d\alpha }}`$ $``$ $`\kappa \lambda ,`$ $`{\displaystyle \frac{d𝒟}{d\alpha }}`$ $``$ $`\kappa \lambda ,`$ $`{\displaystyle \frac{d𝒬}{d\alpha }}`$ $``$ $`\sqrt{𝒬}\kappa \lambda +\lambda ^2.`$ The solution at large $`\alpha `$ would be $`R\kappa \lambda \alpha +R_0`$ and $`𝒟\kappa \lambda \alpha +D_0`$, while $`𝒬`$ is asymptotically given through the implicit equation $$\sqrt{𝒬}\frac{1}{2}\kappa \lambda \alpha +\frac{\lambda }{\kappa }\mathrm{ln}(\sqrt{𝒬}\kappa +\lambda )+\frac{1}{2}\kappa \lambda \kappa _0.$$ (25) Here $`R_0`$, $`D_0`$, and $`\kappa _0`$ are integration constants. Hence, asymptotically, $`\sqrt{𝒬}\frac{1}{2}\kappa \lambda \alpha `$ which is incompatible with the requirement of good generalization $`/\sqrt{𝒬}1`$. Thus the algorithm will not converge, if $`\lambda `$ is kept fixed. The question arises, however, whether a simple extension of the algorithm may not overcome the Odyssean dilemma hinted at in the beginning of this section. We hence suggest to tune the parameter $`\lambda `$ such that it is large enough at small $`\alpha `$ to overcome the pre-asymptotic conditions and it tends to zero at large $`\alpha `$ in order to approach asymptotically the perceptron rule. As can be seen on Fig. 1 this procedure seems successful. Since the situation is now much more complicated we shall not try to solve the general asymptotic problem, as we did in the case of non-structured data, but we shall limit ourselves to prove that robust solutions exist. For this we start with the following ansatz: $`\lambda `$ $`=`$ $`\lambda _0\alpha ^r,`$ (26) $`𝒬`$ $`=`$ $`c^2\alpha ^{2q},`$ (27) $`ϵ_g`$ $`=`$ $`a\alpha ^p,`$ (28) $`\omega `$ $`=`$ $`b\alpha ^s.`$ (29) with $`\omega `$ defined via (20). The asymptotic equations obtained from the flow equations (17)-(19) assuming $`pqr>s0`$ are of the form: $`2\sqrt{𝒬}{\displaystyle \frac{dϵ_g}{d\alpha }}`$ $``$ $`A_{11}\lambda +{\displaystyle \frac{A_{12}}{\sqrt{𝒬}}}+A_2ϵ_g,`$ (30) $`2\sqrt{𝒬}\mathrm{sin}{\displaystyle \frac{z}{2}}ϵ_g{\displaystyle \frac{d\omega }{d\alpha }}`$ $``$ $`B_{11}\lambda +{\displaystyle \frac{B_{12}}{\sqrt{𝒬}}}+B_2ϵ_g,`$ (31) $`{\displaystyle \frac{d\sqrt{𝒬}}{d\alpha }}`$ $``$ $`C_0\lambda +C_1ϵ_g,`$ (32) Here the coefficients $`A_\gamma ,B_\gamma ,C_\gamma `$ are function of $`m,z,L`$ and of $`\omega `$ (the explicit expressions obtained by Maple are given in Appendix, section 5.2). It is easy to see that an asymptotic solution can exist for: $$p=q=r=1/2,s=0,$$ (33) which is therefore compatible with the assumptions used to derive the asymptotic equations (30)-(32). Then $`a,b,c`$ are obtained as function of $`m,\eta ,L`$ for given $`\lambda _0`$, with some restrictions on the latter (notice that the coefficients $`A_\gamma ,B_\gamma ,C_\gamma `$ depend nonlinearly on $`\omega `$, hence on $`b`$). For illustration, we show in Fig. 2 the values of $`a,b`$ and $`c`$ as function of $`\lambda _0`$ for $`L=10,m=1`$ and two values of the data-teacher overlap $`\eta `$. Notice that there is no asymptotic solution for $`\lambda _0`$ below $`0.2`$. In Fig. 3 we show the solution of the full equations (17)-(19) – compare also with Fig. 1 – which can be seen to approach the asymptotic solution (26)-(33). The solutions are robust in the sense that for all $`m,\eta ,L`$ there exists a large region of $`\lambda _0`$ leading to convergence according to (33). Notice, however, that in the pre-asymptotic region similar phenomena to the non-structured data case seem to take place: the flow is divided by a separatrix defined by a $`\lambda _{0,c}`$ (the MC simulation presents the same effect, see Fig. 1). We have thus shown that the simple decrease of $`\lambda `$ as $`1/\sqrt{\alpha }`$ provides convergence to asymptotic perfect generalization with the power $`1/2`$. Alternatively, one can decrease $`\lambda `$ as $`1/\sqrt{𝒬}`$, or as $`e(\alpha )`$, where $$e(\alpha )=\frac{1}{\alpha }\underset{q=1}{\overset{\alpha N/L}{}}e_q$$ (34) using the running “observed error” $`e_q`$ (2) (this is in a sense the most natural choice, since the student only applies its observations). Again the algorithm is stable against noise or a further dilution of the information introduced by taking $`w<1`$. See Fig. 1. ## 4 Summary and Discussion In the present paper we have investigated the performance of the AR-Hebb-algorithm introduced in in the case where the input patterns are structured. The pattern statistics is characterized by the anisotropy vector $`mC`$ and performance of the learning rule depends on $`m`$ and on the overlap $`\eta `$ between the anisotropy vector and the vector $`B`$ that defines the rule – apart from the parameters $`\lambda `$ and $`L`$ which characterize the AR-Hebb-algorithm. As for usual Hebb learning, a tuning of learning parameters is required to achieve good generalization for the classification of structured patterns. Given $`L`$, the only free parameter of the algorithm is $`\lambda `$, and tuning of $`\lambda `$ may proceed in various ways. For instance, one may scale $`\lambda `$ either with $`\alpha `$, i. e., with the number of input-output pairs presented, or with the self-overlap $`𝒬`$, or with the empirical error-rate $`e_q`$. Our analysis reveals that the scaling $`\lambda \alpha ^{1/2}`$, which according to that analysis is equivalent to the scalings $`\lambda 𝒬^{1/2}`$, or $`\lambda e(\alpha )`$, leads to asymptotically perfect generalization. The behaviour is robust in the sense that the prefactor $`\lambda _0`$ may be varied over a wide range without changing the asymptotic scaling of the generalization error. In this sense the tuning required to obtain a working algorithm is not fine-tuning. The only requirement for obtaining good generalization is that $`\lambda _0`$ in (26) exceeds a certain minimum value, $`\lambda _{0,c}^{asympt.}`$. This behaviour is reminiscent of the fact that a minimum value of $`\lambda `$ was also required in the case of unstructured data. In that case, however, the reason was entirely related to pre-asymptotic behaviour related to the fixed-point structure of the flow equations, whereas the above analysis is restricted to the asymptotic domain. From the numerical solution of the full flow-equations and from simulations, we see some empirical evidence that a non-trivial fixed point structure governing the pre-asymptotic behaviour in analogy to what has been found in is present also in the case studied here. As the present dynamical problem is three-dimensional instead two-dimensional, however, the consequences of this might be suspected to be less severe. For instance, a fixed-point with stable and unstable directions in three dimensions does not necessarily produce a separatrix as in the two-dimensional case. However, the projection onto the $`ϵ_g`$-$`\sqrt{𝒬}`$ plane shows a separatrix and hence a $`\lambda _{0,c}`$, as in the unstructured data case (see Fig. 3), with $`\lambda _{0,c}>\lambda _{0,c}^{asympt.}`$. Unlike in the two-dimensional case with unstructured patterns, we have so far not found any evidence of non-universal behaviour of the generalization curve. Whether this is intrinsically related to the different role fixed points appear to play in the present case, we do at present not know. Exceptional behaviour appears for $`\eta =0`$ which in the student-teacher scenario, however, is equivalent to the unstructured case. Acknowledgments: This project was initiated during the Seminar ‘Statistical Physics of Neural Networks’ in Dresden, March 1999. The authors would like to thank the Max Planck Institut für Physik Komplexer Systeme in Dresden for hospitality and financial support and the participants to the workshop for interesting discussions. ## 5 Appendix ### 5.1 Expectations values The expectations values $`A_.,S_.`$ in (17)-(19) are: $`A_{JT}`$ $`=`$ $`m\eta \phi (m{\displaystyle \frac{𝒟}{\sqrt{𝒬}}})+\sqrt{{\displaystyle \frac{2}{\pi }}}{\displaystyle \frac{}{\sqrt{𝒬}}}\mathrm{e}^{\frac{m^2𝒟^2}{2𝒬}}`$ (35) $`=`$ $`m\mathrm{cos}z\phi (m\mathrm{cos}y)+\sqrt{{\displaystyle \frac{2}{\pi }}}\mathrm{cos}x\mathrm{e}^{\frac{1}{2}m^2\mathrm{cos}^2y},`$ $`A_{TT}`$ $`=`$ $`m\eta \phi (m\eta )+\sqrt{{\displaystyle \frac{2}{\pi }}}\mathrm{e}^{\frac{m^2\eta ^2}{2}}`$ (36) $`=`$ $`m\mathrm{cos}z\phi (m\mathrm{cos}z)+\sqrt{{\displaystyle \frac{2}{\pi }}}\mathrm{e}^{\frac{1}{2}m^2\mathrm{cos}^2z},`$ $`S_{JT}`$ $`=`$ $`1+\phi (m{\displaystyle \frac{𝒟}{\sqrt{𝒬}}})\phi (m\eta )4G({\displaystyle \frac{}{\sqrt{𝒬}}},{\displaystyle \frac{𝒟}{\sqrt{𝒬}}},\eta )`$ (37) $`=`$ $`1+\phi (m\mathrm{cos}y)\phi (m\mathrm{cos}z)4G(\mathrm{cos}x,\mathrm{cos}y,\mathrm{cos}z),`$ $`A_{JJ}`$ $`=`$ $`m{\displaystyle \frac{𝒟}{\sqrt{𝒬}}}\phi (m{\displaystyle \frac{𝒟}{\sqrt{𝒬}}})+\sqrt{{\displaystyle \frac{2}{\pi }}}\mathrm{e}^{\frac{m^2𝒟^2}{2𝒬}}`$ (38) $`=`$ $`m\mathrm{cos}y\phi (m\mathrm{cos}y)+\sqrt{{\displaystyle \frac{2}{\pi }}}\mathrm{e}^{\frac{1}{2}m^2\mathrm{cos}^2y},`$ $`A_{TJ}`$ $`=`$ $`m{\displaystyle \frac{𝒟}{\sqrt{𝒬}}}\phi (m\eta )+\sqrt{{\displaystyle \frac{2}{\pi }}}{\displaystyle \frac{}{\sqrt{𝒬}}}\mathrm{e}^{\frac{m^2\eta ^2}{2}}`$ (39) $`=`$ $`m\mathrm{cos}y\phi (m\mathrm{cos}z)+\sqrt{{\displaystyle \frac{2}{\pi }}}\mathrm{cos}x\mathrm{e}^{\frac{1}{2}m^2\mathrm{cos}^2z},`$ $`A_{JC}`$ $`=`$ $`m\phi (m{\displaystyle \frac{𝒟}{\sqrt{𝒬}}})+\sqrt{{\displaystyle \frac{2}{\pi }}}{\displaystyle \frac{𝒟}{\sqrt{𝒬}}}\mathrm{e}^{\frac{m^2𝒟^2}{2𝒬}}`$ (40) $`=`$ $`m\phi (m\mathrm{cos}y)+\sqrt{{\displaystyle \frac{2}{\pi }}}\mathrm{cos}y\mathrm{e}^{\frac{1}{2}m^2\mathrm{cos}^2y},`$ $`A_{TC}`$ $`=`$ $`m\phi (m\eta )+\sqrt{{\displaystyle \frac{2}{\pi }}}\eta \mathrm{e}^{\frac{m^2\eta ^2}{2}}`$ (41) $`=`$ $`m\phi (m\mathrm{cos}z)+\sqrt{{\displaystyle \frac{2}{\pi }}}\mathrm{cos}z\mathrm{e}^{\frac{1}{2}m^2\mathrm{cos}^2z},`$ where $`G({\displaystyle \frac{}{\sqrt{𝒬}}},{\displaystyle \frac{𝒟}{\sqrt{𝒬}}},\eta )`$ $`=`$ $`{\displaystyle \frac{1}{2}}{\displaystyle _{\mathrm{}}^{m\frac{𝒟}{\sqrt{𝒬}}}}{\displaystyle \frac{dt}{\sqrt{2\pi }}}\mathrm{e}^{\frac{1}{2}t^2}\left(1+\phi \left({\displaystyle \frac{t\frac{}{\sqrt{𝒬}}m\eta }{\sqrt{1\frac{^2}{𝒬}}}}\right)\right)`$ (42) $`=`$ $`{\displaystyle \frac{1}{2}}{\displaystyle _{\mathrm{}}^{m\mathrm{cos}y}}{\displaystyle \frac{dt}{\sqrt{2\pi }}}\mathrm{e}^{\frac{1}{2}t^2}\left(1+\phi \left({\displaystyle \frac{t\mathrm{cos}xm\mathrm{cos}z)}{\mathrm{sin}x}}\right)\right)`$ $`\phi (x)`$ $`=`$ $`\mathrm{erf}(x/\sqrt{2})`$ (43) with erf the error function. ### 5.2 Asymptotic coefficients The Maple expressions for the coefficients $`A_\gamma ,B_\gamma ,C_\gamma `$ in (30)-(32) are: $`u`$ $`=`$ $`{\displaystyle \frac{1}{2}}m\mathrm{cos}(z)\sqrt{2}`$ (44) $`v`$ $`=`$ $`m\mathrm{sin}({\displaystyle \frac{1}{2}}z)\sqrt{2}`$ (45) $`g(v\omega )`$ $`=`$ $`v\omega \mathrm{erf}(v\omega )`$ (46) $`f(v\omega )`$ $`=`$ $`v\omega \mathrm{erf}(v\omega )+{\displaystyle \frac{e^{v^2\omega ^2}}{\sqrt{\pi }}}`$ (47) $`A_{11}`$ $`=`$ $`4{\displaystyle \frac{\mathrm{sin}(\frac{1}{2}z)\mathrm{erf}(u)\omega }{\pi }}`$ (48) $`A_{12}`$ $`=`$ $`{\displaystyle \frac{e^{u^2}f(v\omega )}{\pi \sqrt{\pi }L}}`$ (49) $`A_2`$ $`=`$ $`{\displaystyle \frac{e^{u^2}(2{\displaystyle \frac{1+2v^2\omega ^2}{L}}+4(1{\displaystyle \frac{1}{L}})\mathrm{erf}(u)f(v\omega ))}{\sqrt{2\pi }}}`$ (50) $`B_{11}`$ $`=`$ $`4{\displaystyle \frac{m\mathrm{sin}(\frac{1}{2}z)^2\mathrm{erf}(u)(\omega ^21+s^2)}{\pi }}`$ (51) $`B_{12}`$ $`=`$ $`{\displaystyle \frac{\mathrm{sin}(\frac{1}{2}z)e^{u^2}\omega f(v\omega )}{\pi ^{3/2}L}}`$ (52) $`B_2`$ $`=`$ $`4{\displaystyle \frac{m\mathrm{sin}(\frac{1}{2}z)^2e^{u^2}(\omega ^21+s^2)({\displaystyle \frac{v\omega }{L}}(1{\displaystyle \frac{1}{L}})\mathrm{erf}(u)f(v\omega ))}{\sqrt{\pi }}}`$ (53) $`C_0`$ $`=`$ $`\sqrt{2}u\mathrm{erf}(u)+{\displaystyle \frac{\sqrt{2}e^{u^2}}{\sqrt{\pi }}}`$ (54) $`C_1`$ $`=`$ $`(1{\displaystyle \frac{1}{L}})e^{u^2}f(v\omega )(\sqrt{\pi }m\mathrm{erf}(u)\mathrm{cos}(z)+\sqrt{2}e^{u^2})`$ (55)
warning/0001/hep-th0001210.html
ar5iv
text
# Contents ## 1 The Bogoliubov renormalization group ### 1.1 Historical introduction #### 1.1.1 The discovery of the renormalization group In 1952-1953 Stückelberg and Petermann discovered<sup>1</sup><sup>1</sup>1 For a more detailed exposition of the RG early history, see our recent reviews . a group of infinitesimal transformations related to finite arbitrariness arising in $`S`$–matrix elements upon elimination of ultraviolet (UV) divergences. These authors used the notion of normalization group as a Lie transformation group generated by differential operators connected with renormalization of the coupling constant $`e`$. In the following year, on the basis of (infinite) Dyson’s renormalization transformations formulated in the regularized form, Gell-Mann and Low derived functional equations (FEs) for the QED propagators in the UV limit. The appendix to this article contained the general solution (obtained by T.D. Lee) of FE for the renormalized transverse photon propagator amplitude $`d(Q^2/\lambda ^2,e^2)`$ ($`\lambda `$ – cutoff defined as a normalization momentum). This solution was used for a qualitative analysis of the quantum electromagnetic interaction behaviour at small distances. Two possibilities, namely, infinite and finite charge renormalizations were pointed out. However, paper paid no attention to the group character of the analysis and of the qualitative results obtained there. The authors missed a chance to establish a connection between their results and the QED perturbation theory and did not discuss the possibility that a ghost pole solution might exist. The decisive step was made by Bogoliubov and the present author in 1955 <sup>2</sup><sup>2</sup>2See also two survey papers published in English in 1956.. Using the group properties of finite Dyson transformations for the coupling constant, fields and Green functions, they derived functional group equations for the renormalized propagators and vertices in QED in the general (i.e., with the electron mass taken into account) case. In the modern notation, the first equation $$\overline{\alpha }(x,y;\alpha )=\overline{\alpha }(\frac{x}{t},\frac{y}{t};\overline{\alpha }(t,y;\alpha ));x=\frac{Q^2}{\mu ^2},y=\frac{m^2}{\mu ^2},$$ (1) is that for the invariant charge (now widely known also as an effective or running coupling) $`\overline{\alpha }=\alpha d(x,y;\alpha =e^2)`$ and the second — $$s(x,y;\alpha )=s(t,y;\alpha )s(\frac{x}{t},\frac{y}{t};\overline{\alpha }(t,y;\alpha ))$$ (2) — for the electron propagator amplitude. These equations obey a remarkable property: the product $`e^2d\overline{\alpha }`$ of the electron charge squared and the photon transverse propagator amplitude enters into both FEs. This product is invariant with respect to finite Dyson’s transformation (as it is stated by Eq.(1)) which now can be written in the form $$R_t:\left\{\mu ^2t\mu ^2,\alpha \overline{\alpha }(t,y;\alpha )\right\}.$$ (3) We called this product invariant charge and introduced the term renormalization group. Let us emphasize that, unlike in Refs., in the Bogoliubov formulation there is no reference to UV divergences and their subtraction or regularization. At the same time, technically, there is no simplification due to the massless nature of the UV asymptotics. Here, the homogeneity of the transfer momentum scale $`Q`$ is explicitly violated by the mass $`m`$. Nevertheless, the symmetry with respect to transformation $`R_t`$ (even though a bit more involved) underlying RG is formulated as an exact property of the solution. This is what we mean when using the term Bogoliubov renormalization group or renormgroup for short. The differential Lie equations for $`\overline{\alpha }`$ and for the electron propagator $$\frac{\overline{\alpha }(x,y;\alpha )}{\mathrm{ln}x}=\beta (\frac{y}{x},\overline{\alpha }(x,y;\alpha ));\frac{s(x,y;\alpha )}{\mathrm{ln}x}=\gamma (\frac{y}{x},\overline{\alpha }(x,y;\alpha ))s(x,y;\alpha ),$$ (4) with $$\beta (y,\alpha )=\frac{\overline{\alpha }(\xi ,y;\alpha )}{\xi },\gamma (y,\alpha )=\frac{s(\xi ,y;\alpha )}{\xi }\text{at}\xi =1$$ (5) were first derived in by differentiating FEs (1) and (2) over $`x`$ at the point $`t=x`$. On the other hand, by differentiating the same equations over $`t`$ one obtains $$X\overline{\alpha }(x,y;\alpha )=0;Xs(x,y;\alpha )=\gamma (y,\alpha )s(x,y;\alpha )$$ (6) with $$X=x_x+y_y\beta (y,\alpha )_\alpha \left(_x/x\right),$$ (7) the Lie infinitesimal operator. #### 1.1.2 Creation of the RG method Another important achievement of consisted in formulating a simple algorithm for improving an approximate perturbative solution by combining it with Lie group equations – for detail, see below Section 1.3. In our adjacent publication this algorithm was effectively used to analyse the UV and infrared (IR) behaviour in QED. In particular, the one-loop UV asymptotics of the photon propagator as well as the IR behavior of the electron propagator in the transverse gauge $$\overline{\alpha }_{\mathrm{rg}}^{(1)}(x;\alpha )=\frac{\alpha }{1\frac{\alpha }{3\pi }\mathrm{ln}x},s(x,y;\alpha )(p^2/m^21)^{3\alpha /2\pi }$$ (8) were derived. At that time, these expressions, summing the leading log’s terms were already known from papers by Landau with collaborators . However, Landau’s approach did not provide a means for constructing subsequent approximations. The simple technique for calculating higher approximations was found only within the new renormgroup method. In the same paper, starting with the next order perturbation expression $`\overline{\alpha }_{\mathrm{pt}}^{(2)}(x;\alpha )`$ containing the $`\alpha ^3\mathrm{ln}x`$ term, we arrived at the second renormgroup approximation (see below Section 1.3.2) $$\overline{\alpha }_{\mathrm{rg}}^{(2)}(x;\alpha )=\frac{\alpha }{1\frac{\alpha }{3\pi }\mathrm{ln}x+\frac{3\alpha }{4\pi }\mathrm{ln}(1\frac{\alpha }{3\pi }\mathrm{ln}x)}$$ (9) which performs infinite summation of the $`\alpha ^2(\alpha \mathrm{ln})^n`$ terms. This two-loop solution for invariant coupling first obtained in contains the nontrivial log–of–log dependence which is now widely known of the “next-to-leading logs” approximation for the running coupling in quantum chromodynamics (QCD) — see, below, Eq.(21). Comparing (9) with (8), one concludes that two-loop correction is essential in the vicinity of the ghost pole at $`x_1=\mathrm{exp}(3\pi /\alpha )`$. This also shows that the RG method is a regular procedure, within which it is easy to estimate the range of applicability of its results. Quite soon, this approach was formulated for the case of QFT with two coupling constants. To the system of FEs for two invariant couplings there corresponds a coupled system of nonlinear differential equations (DEs). The last was used to study the UV behavior of the $`\pi N`$ interaction at the one-loop level. Thus, in Refs. and RG was directly connected with practical computations of the UV and IR asymptotics. Since then, this technique, the renormalization group method (RGM)<sup>3</sup><sup>3</sup>3Being summarized in the special chapter of the first edition of monograph ., has become the sole means of asymptotic analysis in local QFT. ### 1.2 The Bogoliubov RG: Symmetry of a solution The RG transformation. Generally, RG can be defined as a continuous one-parameter group of specific transformations of a partial solution (or the solution characteristic) of a problem, a solution that is fixed by a boundary condition. The RG transformation involves boundary condition parameters and corresponds to some change in the way of imposing this condition. For illustration, imagine one-argument solution characteristic $`f(x)`$ that has to be specified by the boundary condition $`f(x_0)=f_0`$. Formally, represent a given characteristic of a partial solution as a function of boundary parameters as well: $`f(x)=f(x,x_0,f_0)`$. This step can be treated as an embedding operation. Without loss of generality $`f`$ can be written in a form of a two-argument function $`F(x/x_0,f_0)`$ with the property $`F(1,\gamma )=\gamma `$. The RG transformation then corresponds to a changeover of the way of parameterization, say from $`\{x_0,f_0\}`$ to $`\{x_1,f_1\}`$ for the same solution. In other words, the $`x`$ argument value, at which the boundary condition is given, can be changed for $`x_1`$ with $`f(x_1)=f_1`$. The equality $`F(x/x_0,f_0)=F(x/x_1,f_1)`$ now reflects the fact that under such a change the form of the function $`F`$ itself is not modified. Noting that $`f_1=F(x_1/x_0,f_0)`$, we get $$F(\xi ,f_0)=F(\xi /t,F(t,f_0));\xi =x/x_0,t=x_1/x_0.$$ The group transformation here is $`\{\xi \xi /t,f_0F(t,f_0)\}`$. The renormgroup transformation for a given solution of some physical problem in the simplest case can now be defined as a simultaneous one-parameter transformation of two variables, say $`x`$ and $`g`$, by $$R_t:\{xx^{}=x/t,gg^{}=\overline{g}(t,g)\},$$ (10) the first being a scaling of a coordinate $`x`$ (or reference point) and the second — a more complicated functional transformation of the solution characteristic. The equation $$\overline{g}(x,g)=\overline{g}(x/t,\overline{g}(t,g))$$ (11) for the transformation function $`\overline{g}`$ provides the group property $`T_{\tau t}=T_\tau T_t`$ of the transformation (10). They are just the RG FEs and transformation for a massless QFT model with one coupling constant $`g`$. In that case $`x=Q^2/\mu ^2`$ is the ratio of a 4-momentum $`Q`$ squared to a “normalization” momentum $`\mu `$ squared and $`g`$, the coupling constant. The RG transformation (10) of a QFT amplitude $`s`$ is of the form (compare with Eq.(2)) $$R_ts(x,g)e^{\mathrm{ln}tX}s(x,g)=s(x/t,\overline{g}(t,g))=z_s^1s(x,g);z_s=s(t,g).$$ (12) Several generalizations are in order. a. “Massive” case. For example, in QFT, if we do not neglect mass $`m`$ of a particle, we have to insert an additional dimensionless argument into the invariant coupling $`\overline{g}`$ which now has to be considered as a function of three variables: $`x=Q^2/\mu ^2,y=m^2/\mu ^2`$, and $`g`$. The presence of a new “mass” argument $`y`$ modifies the group transformation (10) and the FE (11) $$R_t:\left\{x^{}=\frac{x}{t},y^{}=\frac{y}{t},g^{}=\overline{g}(t,y;g)\right\};\overline{g}(x,y;g)=\overline{g}(\frac{x}{t},\frac{y}{t};\overline{g}(t,y;g)).$$ (13) Here, it is important that the new parameter $`y`$ (which, physically, should be close to the $`x`$ variable, as it scales similarly) enters also into the transformation law of $`g`$ . If the considered QFT model, like QCD, contains several masses, there will be several mass arguments $`y\{y\}y_1,y_2,\mathrm{},y_n.`$ b. Multi-coupling case. A more involved generalization corresponds to transition to the case with several coupling constants: $`g\{g\}=g_1,\mathrm{},g_k`$. Here, there arise a “family” of effective couplings $$\overline{g}\{\overline{g}\},\overline{g}_i=\overline{g}_i(x,y;\{g\});i=1,2,\mathrm{},k,$$ (14) satisfying the system of coupled functional equations $$\overline{g}_i(x,y;\{g\})=\overline{g}_i(x/t,y/t;\{\overline{g}(t,y;\{g\})\}).$$ (15) The RG transformation now is $$R_t:\left\{xx/t,yy/t,\{g\}\{g(t)\}\right\},g_i(t)=\overline{g}_i(t,y;\{g\}).$$ (16) ### 1.3 Renorm-group method #### 1.3.1 The algorithm The idea of the approximate solution marriage to group symmetry can be realized with the help of group DEs. If we define $`\beta `$ and $`\gamma `$ (the so-called “generators” on physical slang) from some approximate solutions and then solve evolutional DEs, we obtain the RG improved solutions that obey the group symmetry and correspond to the approximate solutions used as an input. Now we can formulate an algorithm of improving an approximate solution. The procedure is given by the following prescription which we illustrate by a massless one–coupling case (4) and (5): Assume some approximate solution $`\overline{g}_{\mathrm{appr}}(x,g),s_{\mathrm{appr}}(x,g)`$ is known. 1. On the basis of Eq.(5) define the beta– and gamma–functions $$\beta (g)\stackrel{\mathrm{def}}{=}\frac{}{\xi }\overline{\mathrm{g}}_{\mathrm{appr}}(\xi ,\mathrm{g})|_{\xi =1};\gamma (\mathrm{g})\stackrel{\mathrm{def}}{=}\frac{}{\xi }\mathrm{s}_{\mathrm{appr}}(\xi ,\mathrm{g})|_{\xi =1}.$$ (17) 2. Integrate the first of Eqs.(4), i.e., construct the function $$f(g)\stackrel{\mathrm{def}}{=}^\mathrm{g}\frac{\mathrm{d}\tau }{\beta (\tau )},$$ (18) 3. Resolve the obtained equation, i.e., $$\overline{g}_{\mathrm{rg}}(x,g)=f^1\{f(g)+\mathrm{ln}x\}.$$ (19) 4. Integrate the second of Eqs.(4) using this expression $`\overline{g}_{\mathrm{rg}}`$ in its r.h.s. to obtain $`s_{\mathrm{rg}}(x,g)`$ in the explicit form. 5. Then, the expressions $`\overline{g}_{\mathrm{rg}}`$ and $`s_{\mathrm{rg}}`$ precisely satisfy the RG symmetry, i.e., they are exact solutions of Eqs.(11) and (12) corresponding to $`\overline{g}_{\mathrm{appr}}`$ and $`s_{\mathrm{appr}}`$ used as an input. #### 1.3.2 Simple illustration For illustration, take the simplest perturbative expression $`\overline{g}_{\mathrm{pt}}^{(1)}=gg^2\beta _1\mathrm{ln}x`$ for $`\overline{g}_{\mathrm{appr}}`$ and $`s_{\mathrm{pt}}^{(1)}=1g\gamma _1\mathrm{ln}x`$. Here, $`\beta (g)=\beta _1g^2,\gamma (g)=\gamma _1g`$ and integration of (4) gives explicit expressions $$\overline{g}_{\mathrm{rg}}^{(1)}(x,g)=\frac{g}{1+g\beta _1\mathrm{ln}x},s_{\mathrm{rg}}^{(1)}(x,g)=\left(\overline{g}(x,g)/g\right)^{\nu _1};\nu _1=\gamma _1/\beta _1,$$ (20) which, on the one hand, exactly satisfy the RG symmetry and, on the other, being expanded in powers of $`g`$, correlate with $`\overline{g}_{\mathrm{pt}}`$ and $`s_{\mathrm{pt}}`$. Now, on the basis of geometric progression (20), let us present the 2-loop perturbative approximation for $`\overline{g}`$ in the form $`\overline{g}_{\mathrm{pt}}^{(2)}=gg^2\beta _1\mathrm{ln}x+g^3(\beta _1^2\mathrm{ln}^2x\beta _2\mathrm{ln}x).`$ By using this expression as an input in Eq.(17), we have $`\beta ^{(2)}(g)=\beta _1g^2\beta _2g^3`$ and then (step 2) $$\beta _1f^{(2)}(z)=^z\frac{d\tau }{\tau ^2+b\tau ^3}=\frac{1}{z}+b\mathrm{ln}\frac{z}{1+bz};b=\frac{\beta _2}{\beta _1}.$$ To make the last step, we have to start with the equation $$f^{(2)}[\overline{g}_{\mathrm{rg}}^{(2)}(x,g)]=f^{(2)}(g)+\beta _1\mathrm{ln}x$$ which is a transcendental one and has no simple explicit solution<sup>4</sup><sup>4</sup>4It can be expressed via special, Lambert, $`W`$-function: $`W(z)\mathrm{exp}^{W(z)}=z`$; see, e.g., Ref... Due to this, one usually resolves this relation approximately. Take into account that the second, logarithmic, contribution to $`f^{(2)}(z)`$ is a small correction to the first one at $`bz1`$. Under this reservation, we can substitute the one-loop RG expression (20) instead of $`\overline{g}_{\mathrm{rg}}^{(2)}`$ into this correction and obtain the explicit “iterative” solution $$\overline{g}_{\mathrm{rg}}^{(2)}=\frac{g}{1+g\beta _1l+g(\beta _2/\beta _1)\mathrm{ln}\left[1+g\beta _1l\right]};l=\mathrm{ln}x.$$ (21) An analogous procedure for $`s_{\mathrm{pt}}^{(2)}=1g\gamma _1\mathrm{ln}x+g^2\left(\gamma _1(\gamma _1+1)\beta _1^2\mathrm{ln}^2x\gamma _2\mathrm{ln}x\right)`$ yields $$s_{\mathrm{rg}}^{(2)}=\frac{S\left(\overline{g}_{\mathrm{rg}}^{(2)}(x,g)\right)}{S(g)}\text{with}S(g)=g^{\nu _1}e^{\nu _2g}\text{and}\nu _2=\frac{\beta _1\gamma _2\gamma _1\beta _2}{\beta _1^2}.$$ (22) These results are interesting from several aspects. * First, being expanded in $`g`$ and $`gl`$ powers, they produce an infinite series containing “leading” and “next-to-leading” UV logarithmic contributions. * Second, they contain a nontrivial analytic dependence $`\mathrm{ln}(1+g\beta _1l)\mathrm{ln}(\mathrm{ln}Q^2)`$ which is absent in the perturbation input. * Third, being compared with the one-loop solution, Eq.(20), they demonstrate an algorithm of subsequent improving of accuracy, i.e., of RGM regularity. #### 1.3.3 RGM usage in QFT As we have seen, QFT perturbation expression of finite order does not obey the RG symmetry. On the other hand, it was shown that the one-loop and two-loop approximations, used as an input for the construction of “generators” $`\beta (g)`$ and $`\gamma (g)`$, yield expressions (20), (21) and (22) that obeys the group symmetry and exactly satisfy FEs (11) and (12). More generally, one can state the following logical structure of the RGM procedure. — Solving group equation(s) for invariant coupling(s) $`\overline{g}_{\mathrm{rg}}(x,g)`$ using some approximate solution $`\overline{g}_{\mathrm{pt}}`$ as an input. — Obtaining RG solutions for some other QFT objects (like vertices and propagator amplitudes) on the basis of the expression(s) for $`\overline{g}_{\mathrm{rg}}`$ just derived. Typically, they satisfy the equation $$XM(x,y,g)=\gamma (y,g)M(x,y,g).$$ (23) General structure of the corresponding solutions has the form $$M(x,y;g)=z_M^1(y,g)(x/y,\overline{g}(x,y;g)).$$ (24) Note that the function $``$ in the r.h.s. depends only on the RG invariants, that is on the first integrals of the RG operator $`X`$ introduced in Eqs.(6) and (7). It satisfies homogeneous partial differential equations (PDEs) $`X=0`$. For the RG invariant objects, like observables, $`z_M=1`$, $`\gamma =0`$. Now we can resume the RGM properties. The RGM is a regular procedure of combining dynamical information (taken from an approximate solution) with the RG symmetry. The essence of RGM is the following: 1) The mathematical tool used in RGM is Lie differential equations. 2) The key element of RGM is possibility of an (approximate) determination of “generators”, like $`\beta (g),\gamma (g)`$, from dynamics. 3) The RGM works effectively in the case when the solution has a singular behaviour. It restores the structure of singularity compatible with the RG symmetry. ## 2 Evolution of Renormalization Group In the 70s and 80s RG ideas have been applied to critical phenomena: spontaneous magnetization, polymerization, percolation, non-coherent radiation transfer, dynamic chaos, and so on. Less sophisticated motivation by Wilson in spin lattice phenomena (than in QFT) made this “explosion” of RG applications possible. ### 2.1 Renormalization Group Evolving #### 2.1.1 Kadanoff–Wilson RG in critical phenomena a. Spin lattice. The so–called renormalization group in critical phenomena is based on the Kadanoff–Wilson procedure referred to as “decimation” or “blocking”. Initially, it emerged from the problem of spin lattice. Imagine a regular (two– or three–dimensional) lattice consisting of $`N^d,d=2,3`$ cites with an ‘elementary step’ $`a`$ between them. Suppose that at every site a spin vector $`𝝈`$ is located. The Hamiltonian, describing the spin interaction between nearest neighbours $$H=k\underset{i}{}𝝈_i𝝈_{i\pm 1}$$ contains $`k`$, the coupling constant. A statistical sum is obtained from the partition function, $`S=<\mathrm{exp}(H/\theta )>_{\mathrm{aver}}.`$ To realize blocking, one has to perform the “spin averaging” over block consisting of $`n^d`$ elementary sites. This step diminishes the number of degree of freedom from $`N^d`$ to $`(N/n)^d`$. It also destroys the small-range properties of a system, in the averaging course some information being lost. However, the long-range physics (like correlation length essential for phase transition) is not affected by it, and we gain simplification of the problem. As a result of this blocking procedure, new effective spins $`𝚺`$ arise in new sites forming a new effective lattice with a step $`na`$. We arrive also at the new effective Hamiltonian $$H_{\mathrm{eff}}=K_n\underset{I}{}𝚺_I𝚺_{I\pm 1}+\mathrm{\Delta }H,$$ with the effective coupling $`K_n`$ between new spins $`𝚺_I`$ of new neighbouring sites; $`K_n`$ has to be defined by the averaging process as a function of $`k`$ and $`n`$. Here, $`\mathrm{\Delta }H`$ contains quartic and higher spin forms which are irrelevant for the IR (long-distance) properties. Due to this, one can drop $`\mathrm{\Delta }H`$ and conclude that the spin averaging leads to an approximate transformation, $$k\underset{i}{}𝝈𝝈K_n\underset{I}{}𝚺𝚺,$$ or, taking into account the “elementary step” change, to $`\left\{ana,kK_n\right\}`$. The latter is the Kadanoff–Wilson transformation. It is convenient to write down the new coupling $`K_n`$ in the form $`K_n=K(1/n,K)`$. Then, the KW transformation reads $$KW_n:\left\{ana,kK_n=K(1/n,k)\right\}.$$ (25) These transformations obey composition law $`KW_nKW_m=KW_{nm}`$ if the relation $$K(x,k)=K(x/t,K(t,k)),x=1/nm,t=1/n.$$ (26) holds. This is very close to RG symmetry. We observe the following points: * The RG symmetry is approximate (due to neglecting $`\mathrm{\Delta }H`$). * The transformations $`KW_n`$ are discrete. * There exist no reverse transformation to $`KW_n`$. * Transformations $`KW_n`$ relate different auxiliary models. Hence, the ‘Kadanoff–Wilson renormalization group‘ (KW–RG) is an approximate and discrete semi-group. For a long–distance (IR limit) physics, however, $`\mathrm{\Delta }(1/n)`$ is small and it is possible to use differential Lie equations<sup>5</sup><sup>5</sup>5In application of these transformations to critical phenomena, the notion of a fixed point is important. Generally, a fixed point is associated with power-type asymptotic behavior. Note here that, contrary to the QFT case considered in Section 1.3.2, in phase transitions we deal with the IR stable point.. b. Polymer theory. In polymer physics, one considers statistical properties of polymer macromolecules which can be imagined as a very long chain of identical elements (with the number of elements $`N`$ as big as $`10^5`$). Molecules are swimming in a solvent and form globulars. This big molecular chain forms a specific pattern resembling the pattern of a random walk. The central problem of the polymer theory is very close to that of a random walk and can be formulated as follows. For a long chain of $`N`$ “steps” (the size of step = $`a`$), one has to find the “chain size” $`R_N`$, the distance between the “start” and the “finish” points (the size a of globular), with the distribution function $`f(\varphi )`$ of angles between the neighboring elements being given. For large $`N`$ values, the molecular size $`R_N`$ follows the power Fleury law $`R_NN^\nu `$ with $`\nu `$, the Fleury index. When $`N`$ is given, $`R_N`$ is a functional of $`f(\varphi )`$ which depends on external conditions (e.g., temperature $`T`$, properties of solvent, etc. ). If $`T`$ grows, $`R_N`$ increases and at some moment globulars touch one another. This is the polymerization process which is very similar to a phase transition phenomenon. The Kadanoff–Wilson blocking ideology has been introduced in physics of polymers by De Gennes . The key idea is a grouping of $`n`$ neigbouring elements of a chain into a new “elementary block”. It leads to the transformation $`\left\{1n;aA_n\right\}`$ which is analogous to one for the spin lattice decimation. This transformation must be specified by a direct calculation which gives an explicit form of $`A_n=\overline{a}(n,a)`$. Here, we have a discrete semi-group. Then, by using the KW–RG technique, one finds the fixed point, obtains the Fleury power law and can calculate its index $`\nu `$. The essential feature of a polymer chain is the impossibility of a self-intersection. This is known as an excluded volume effect in the random walk problem. Generally, the excluded volume effect yields some complications. However, inside the alternative, the QFT RG approach to polymers , it can be treated rather simply by introducing an one more argument which is similar to finite length $`L`$ in the transfer problem or particle mass $`m`$ in QFT. Besides polymers, the KW–RG technique has been used in some fields of physics, like percolation, non-coherent radiation transfer , dynamical chaos and some others. #### 2.1.2 Bogoliubov symmetry outside QFT Meanwhile, the original QFT–RG approach proliferated into some other parts of theoretical physics. In the late 50s, it was used for summation of Coulomb singularities in Bogoliubov’s theory of superconductivity based on the Fröhlich electron–phonon interaction. Twenty years later it was used in the theory of turbulence. a. Turbulence. To formulate the turbulence problem in terms of RG, one has to perform the following steps : 1. Introduce the generating functional for correlation functions. 2. Write down the path integral representation for this functional. 3. By changing the functional integration variable, find the equivalence of the statistical system to some quantum field theory model. 4. Construct the system of Schwinger–Dyson equations for this equivalent QFT model. 5. Perform the finite renormalization procedure and derive the RG equations. Here, the reparameterization degree of freedom, physically corresponds to a change of long wave-length cutoff which is built-in into the definition of a few effective parameters. b. Weak shock wave. Another example can be taken from hydrodynamics. Consider a weak shock wave in the one-dimensional case of a large distance $`l`$ from the starting (implosion) point. The dependence of velocity $`v`$ of a matter as a function of $`l`$ at a given moment of time $`t`$ has a simple triangular shape and can be described by the expression $$v(l)=\frac{l}{L}V\text{at}lL;=0\text{for}l>L,$$ where $`L=L(t)`$ is the front position and $`V=v(L)`$ – the front velocity. They are functions of time. In the absence of viscosity, the “conservation law” $`LV=\mathrm{Const}.`$ holds. Due to this, they can be treated as functions of the front wave position $`Lx,V=V(x)`$ as well. If the physical situation is homogeneous, then the front velocity $`V(x)`$ should be considered as a function of only two additional relevant arguments – its own value $`V_0=V(x_0)`$ at some precedent point $`(x_0<x)`$ and of the $`x_0`$ coordinate. In can be written down in the form : $`V(x)=G(x/x_0,V_0)`$. If we pick up three points $`x_0`$, $`x_1`$ and $`x_2`$ (for details, see Refs.), then the initial condition may be given either at $`x_0`$ or $`x_1`$. Thus, we obtain the FE equation equivalent to (11) $$V_2=G(x_2/x_0,V_0)=G(x_2/x_1,V_1)=G(x_2/x_1,G(x_1/x_0,V_0)).$$ c. One-dimensional transfer. A similar argument has been done by Mnatzakanian in the transfer problem at one dimension. Imagine a half–space filled with a homogeneous medium on the surface of which some flow (of radiation or particles) with intensity $`g_0`$ falls from the vacuum half–space. Follow the flow as it moves inwards the medium at the distance $`l`$ from the boundary. Due to homogeneity along the $`l`$ coordinate, the intensity of the penetrated flow $`g(l)`$ depends on two essential arguments, $`g(l)=G(l,g_0)`$. The values of the flow at three different points $`g_0`$ (on the boundary), $`g_1`$ and $`g_2>g_1`$ can be connected with each other by the transitivity relations, $`g_1=G(\lambda ,g_0),g_2=G(\lambda +l,g_0)=G(l,g_1),`$ which lead to the FE $$G(l,g)=G(l\lambda ,G(\lambda ,g)).$$ (27) Performing a logarithmic change of variables $`l=\mathrm{ln}x,\lambda =\mathrm{ln}t,G(l,g)=\overline{g}(x,g)`$, we see that (27) is equivalent to (11). Consider now intensity of a reverse flow, that is total amount of particles at the point $`l`$ moving in the backward direction. It is completely defined by $`g_0`$ and can be written down as $`R(l,g_0)`$. This function can be represented in the form $`R(l,g)=R_0(g)N(l,g)`$ with $`R_0R(0,g)`$ and function $`N`$ “normalized” on the boundary $`N(0,g)=1`$. Playing the same game with transitivity, we arrive at FE $$N(l,g)=Z(l,g)N(l\lambda ,G(l,G(\lambda ,g));Z=R_0(g_1)/R_0(g)$$ (28) related to Eq.(12) by logarithmic change of variables. One can refer to (27) and (28) as to the additive version of RG FEs and to previous equations of Section 1, like (11), (12) and (13) as to the multiplicative one. The transfer problem admits a modification connected with discrete inhomogeneity: imagine the case of two different kinds of homogeneous materials separated by the inner boundary surface at $`l=L`$. The point of breaking $`l=L`$ may correspond to the boundary with empty space, and resulting equation is equivalent to Eq.(13). One more generalization is related to “multiplication” of argument $`g`$ as expressed by Eq.(14). Physically, this relates to the case of radiation on different frequencies $`\omega _i,i=1,2,\mathrm{}k`$ (or particles of different energies or of different types). Take the case of $`k=2`$ and suppose that the material of the medium has such properties that the transfer processes of the two flows are not independent. In this case, the characteristic functions of these flows $`G`$ and $`H`$ are dependent on both the boundary values $`g_0`$ and $`h_0`$ and can be taken as functions $`g(l)=G(l,g_0,h_0),h(l)=H(l,g_0,h_0)`$. After a group operation $`ll\lambda `$, we arrive at a coupled set of functional equations $$G(l+\lambda ,g,h)=G(l,g_\lambda ,h_\lambda ),H(l+\lambda ,g,h)=H(l,g_\lambda ,h_\lambda );g_\lambda G(\lambda ,g,h),h_\lambda H(\lambda ,g,h)$$ which is just an additive version of system (15) at $`k=2`$. Now we can make the important conclusion that a common property yielding functional group equations is just the transitivity property of some physical quantity with respect to the way of giving its boundary or initial value. Hence, the RG symmetry is not a symmetry of equations but a symmetry of solution, that is of equations and boundary conditions considered as a whole. ### 2.2 Difference between Bogoliubov RG and KW–RG As we have mentioned above, the RG ideas expanded in diverse fields of physics in two different ways: * via direct analogy with the Kadanoff–Wilson construction (averaging over some set of degrees of freedom) in polymers, non-coherent transfer and percolation, i.e., constructing a set of models for a given physical problem. * via finding an exact RG symmetry by proof of the equivalence with a QFT model (e.g., in turbulence ), plasma turbulence ) or by some other reasoning (like in a transfer problem). To the question Are there different renormalization groups? the answer is positive: 1. In QFT and some simple macroscopic examples, RG symmetry is an exact symmetry of the solution formulated in its natural variables. 2. In turbulence, continuous spin-field models and some others, it is a symmetry of an equivalent QFT model. 3. In polymers, percolation, etc. , (with KW blocking), the RG transformation is a transformation between different auxiliary models (specially constructed for this purpose) of a given system. As we have shown, there is no essential difference in the mathematical formulation. There exists, however, a profound difference in physics: — In the cases 1 and 2 (as well as in some macroscopic examples), the RG is an exact symmetry of a solution. — In the Kadanoff–Wilson type problem (spin lattice, polymers, etc. ), one has to construct a set $``$ of models $`M_i`$. The KW–RG transformation $$KW_nM_i=M_{ni},\text{with integer}n$$ (29) is acting inside a set of models. ### 2.3 Functional self-similarity The RG transformations have close connection with the concept of self-similarity. The self-similarity transformations for problems formulated by nonlinear PDEs are well known since the last century, mainly in dynamics of liquids and gases. They are one parameter $`\lambda `$ transformations defined as a simultaneous power scaling of independent variables $`z=\{x,t,\mathrm{}\}`$, solutions $`f_k(z)`$ and other functions $`V_i(z)`$ (like external force) $$S_\lambda :\left\{x^{}=x\lambda ,t^{}=t\lambda ^a,f_k^{}=\lambda ^{\phi _k}f_k,V_i^{}=\lambda ^{\nu _i}V_i\right\}$$ entering into the equations. To emphasize their power structure, we use a term power self-similarity = PS. According to Zel’dovich and Barenblatt, the PS can be classified as: a/ PS of the 1st kind with all indices $`a,\mathrm{}\phi ,\nu ,\mathrm{}`$ being integers or rational (Rational PS) that are usually found from the theory of dimensions; b/ PS of the 2nd kind with irrational indices (Fractal PS) which should be defined from dynamics. To relate RG with PS, turn to the renormgroup FE $`\overline{g}(xt,g)=\overline{g}(x,\overline{g}(t,g))`$. Its general solution is known; it depends on an arbitrary function of one argument – see Eq.(19). However, at the moment, we are interested in a special solution linear in the second argument: $`\overline{g}(x,g)=gX(x).`$ The function $`X(x)`$ should satisfy the equation $`X(xt)=X(x)X(t)`$ with the solution $`X(x)=x^\nu `$. Hence, $`\overline{g}(x,t)=gx^\nu `$. This means that in our special case, linear in $`g`$, the RG transformation (10) is reduced to the PS transformation, $$R_tS_t:\{x^{}=xt^1,g^{}=gt^\nu \}.$$ (30) Generally, in RG, instead of a power law, we have an arbitrary functional dependence. Thus, one can consider transformations (10), (13) and (16) as functional generalizations of usual (i.e., power) self-similarity transformations. Hence, it is natural to refer to them as to the transformations of functional scaling or functional (self)similarity (FS) rather than to RG-transformations. In short, $$\mathrm{RG}\mathrm{FS},$$ with FS standing for Functional Similarity<sup>6</sup><sup>6</sup>6This notion was first mentioned in and formally introduced in the beginning of 80s.. Now we can answer the question on the physical meaning of the symmetry underlying FS and the Bogoliubov renormgroup. As we have mentioned, it is not a symmetry of a physical system or of equation(s) of the problem at hand, but a symmetry of a solution considered as a function of the relevant physical variables and suitable boundary parameters. A symmetry like that can be related, in particular, to the invariance of a physical quantity described by this solution with respect to the way in which the boundary conditions are imposed. The changing of this way constitutes a group operation in the sense that the group composition law is related to the transitivity property of such changes. Homogeneity is an important feature of a physical system under consideration. However, homogeneity can be violated in a discrete manner. Imagine that such a discrete violation is connected with a certain value of $`x`$, say, $`x=y`$. Here, RG transformation with the canonical parameter $`t`$ has the form (13). The symmetry connected with FS is a very simple and frequently encountered property of physical solutions. It can easily be “discovered” in numerous problems of theoretical physics like classical mechanics, transfer theory, classical hydrodynamics, and so on – see, above, Section 2.1.2. ## 3 Symmetry of solution in Mathematical Physics ### 3.1 Constructing RG-symmetries and their use From the discussion in Sections 1.1 and 1.2 it follows that FS transformation in QFT is the scaling transformation of an independent variable $`x`$ (and, possibly, the parameter $`y`$) accompanied by a functional transformation of the solution characteristic $`g`$. It is introduced by means of either finite transformations (10), (13) and (16) or the infinitesimal operator (7). Hence, the symmetry of a solution, i.e., FS symmetry, is commonly understood in QFT as the Lie point symmetry of a one-parameter transformation group defined by the operator of the (7)–type. Now, we are interested in getting answers to the following questions: * is it possible to extend the notion of RG symmetry (RGS) and generalize the form of RGS implementation that may differ from that given by (7) ? — and if ”yes”, * is it possible to create a regular algorithm of finding these symmetries ? The answer is positive to both the questions, and below we demonstrate the regular algorithm of constructing RGS in mathematical physics that up to now has been devised only for boundary value problem (BVP) for the (system of) differential equation(s) which we shall refer to as basic equations (BEs). The point is that these models can be analyzed by methods of Lie group analysis which employ infinitesimal group transformations instead of the finite one. The general idea of the algorithm is to find a specific renormgroup manifold $``$ that contains the desired solution of BVP. Then construction of a RGS, that leaves this solution unaltered, is performed by standard methods of a group analysis of DEs. The regular algorithm of constructing RGS (and their application) can be formulated in a form of a scheme<sup>7</sup><sup>7</sup>7In the present form this scheme was described in . One can find there historical comments and references on the pioneering publications. which comprises a few steps. It is illustrated in Figure 1. I. First of all, a specific renormgroup manifold $``$ for the given BVP should be constructed which is identified below with a system of the $`k`$th-order DEs $$F_\sigma (z,u,u_{(1)},\mathrm{},u_{(k)})=0,\sigma =1,\mathrm{},s.$$ (31) In (31) and what follows we use terminology of group analysis and the notation of differential algebra. In contrast with the mathematical analysis, where we usually deal with functions $`u^\alpha ,\alpha =1,\mathrm{},m`$ of independent variables $`x^i,i=1,\mathrm{},n`$ and derivatives $`u_i^\alpha (x)u^\alpha /x^i,u_{ij}^\alpha (x)^2u^\alpha /x^ix^j,\mathrm{}`$ that are also considered as functions of $`x`$, in differential algebra we treat $`u^\alpha ,u_i^\alpha ,u_{ij}^\alpha ,\mathrm{}`$ as variables as well. Therefore, in differential algebra we deal with an infinite number of variables $$x=\{x^i\},u=\{u^\alpha \},u_{(1)}=\{u_i^\alpha \},u_{(2)}=\{u_{i_1i_2}^\alpha \},\mathrm{};(i,i_1,\mathrm{}=1,\mathrm{},n),$$ (32) where $`x^i`$ are called independent variables, $`u^\alpha `$ dependent variables and $`u_{(1)},u_{(2)},\mathrm{}`$ derivatives. A locally analytic function $`f(x,u,u_{(1)},\mathrm{},u_{(k)})`$ of variables (32), with the highest $`k`$th-order derivative involved, is called a differential function of order $`k`$. The set of all differential functions of a given order form a space of differential functions $`𝒜`$, the universal space of modern group analysis . A particular way of realization of the first step can hardly be described uniquely, as it depends on both a form of basic equations and a boundary condition; generally, $``$ does not coincide with BEs. We indicate here a few feasible routines for this step. * One can use an extension of the space of variables involved in group transformations. These variables, e.g., may be parameters, $`p=\{p^j\},j=1,\mathrm{},l`$ entering into a solution via the equations and/or boundary conditions. Adding parameters $`p`$ to the list of independent variables $`z=\{x,p\}`$ we treat BEs in this extended space as $``$ (31). Similarly, one can extend the space of differential variables by treating derivatives with respect to $`p`$ as additional differential variables. * Another possibility employs reformulating of boundary conditions in terms of embedding equations or differential constraints which are then combined with BEs. The key idea here is to treat the solution of BVP as an analytic function of independent variables and boundary parameters $`b=\{x_0^i,u_0^\alpha \}`$ as well. Differentiation with respect to these parameters gives additional DEs (embedding equations) that, together with BEs, form $``$. In some cases, while calculating Lie point RGS, the role of embedding equations can be played by differential constraints (for details see ) that come from an invariance condition for BEs with respect to the Lie-Bäcklund <sup>8</sup><sup>8</sup>8We use here the terminology adopted in Russian literature . This symmetry is also known as generalized or higher-order symmetry . symmetry group. * In the case when BEs contain a small parameter $`\alpha `$, the desired $``$ can be obtained by simplification of these equations and use of “perturbation methods of group analysis” (see Vol.3, Chapter 2, p.31 in ). The main idea here is to consider a simplified ($`\alpha =0`$) model, which admits a wider symmetry group (see examples in the Section 4.2 below) in comparison with the case $`\alpha 0`$. When we take the contributions from small $`\alpha `$ into account, this symmetry is inherited by BEs, which results in the additional terms, corrections in powers of $`\alpha `$, in the RGS generator. II. The next step consists in calculating the most general symmetry group $`𝒢`$ that leaves the manifold $``$ unaltered. The term “symmetry group”, as used in the classical group analysis, means the property of the system (31) to admit a local Lie group of point transformations in the space $`𝒜`$. The Lie algorithm of finding such symmetries consists in constructing tangent vector fields defined by the operator $$X=\xi ^i_{x^i}+\eta ^\alpha _{u^\alpha },\xi ^i,\eta ^\alpha 𝒜,$$ (33) with the coordinates, $`\xi ^i,\eta ^\alpha `$ that are functions of group variables and have to be determined by a system of equations $$XF_{\sigma }^{}{}_{|(\text{31})}{}^{}=0,\sigma =1,\mathrm{},s,$$ (34) that follow from the invariance of $``$. Here $`X`$ is extended<sup>9</sup><sup>9</sup>9The extending of generators to the derivatives employs the prolongation formulas and is a regular procedure in group analysis (see, e.g. ). to all derivatives involved in $`F_\sigma `$ and the symbol $`_{(\text{31})}`$ means calculated on the frame (31). A system of linear homogeneous PDEs (34) for coordinates $`\xi ^i,\eta ^\alpha `$, known as determining equations, is an overdetermined system as a rule. The solution of Eqs.(34) define a set of infinitesimal operators (33) (also known as group generators), which correspond to the admitted vector field and form a Lie algebra. In the case that the general element of this algebra $$X=\underset{j}{}A^jX_j,$$ (35) where $`A^j`$ are arbitrary constants, contains finite number of operators, $`1jl`$, the group is called finite-dimensional (or simply finite) with the dimension $`l`$; otherwise, for unlimited $`j`$ or in the case that coordinates $`\xi ^i`$, $`\eta ^\alpha `$ depend upon arbitrary functions of group variables, the group is called infinite. The use of the infinitesimal criterion (34) for calculating the symmetry groups makes the whole procedure algorithmic and can be carried out not only “by hand” but using the symbolic packages of the computer algebra (see, e.g., Vol.3 in ) as well. In modern group analysis, different modifications of the classical Lie scheme are in use (see and references therein). The generator (33) of the group $`𝒢`$ is equivalent to the canonical Lie–Bäcklund operator $$Y=\kappa ^\alpha _{u^\alpha },\kappa ^\alpha \eta ^\alpha \xi ^iu_i^\alpha ,$$ (36) that is known as a canonical representation of $`X`$ and plays an essential role in RGS constructing. However, the group defined by the generators (33) and (36) cannot yet be referred to as a renormgroup, as it is not related to a partial BVP solution of interest. III. To obtain RGS, the restriction of the group $`𝒢`$ on particular BVP solution should be made which forms the third step. Mathematically, this procedure appears as checking the vanishing condition for the linear combination of coordinates $`\kappa _j^\alpha `$ of the canonical operator equivalent to (35) on a particular approximate (or exact) BVP solution $`U^\alpha (z)`$ $$\left\{\underset{j}{}A^j\kappa _j^\alpha \underset{j}{}A^j\left(\eta _j^\alpha \xi _j^iu_i^\alpha \right)\right\}_{|u^\alpha =U^\alpha \left(z\right)}=0.$$ (37) Evaluating (37) on a particular BVP solution $`U^\alpha (z)`$ transforms the system of DEs for group invariants into algebraic relations<sup>10</sup><sup>10</sup>10Similar relations were discussed in , Chapter 8, when constructing invariant solutions for the Cauchy problem for a quasi-linear system of first order PDEs.. Firstly, it gives relations between $`A^j`$ thus “combining” different coordinates of group generators $`X_j`$ admitted by the $``$ (31). Secondly, it eliminates (partially or entirely) the arbitrariness that may appear in coordinates $`\xi ^i`$, $`\eta ^\alpha `$ in the case of an infinite group $`𝒢`$. In terms of the “classic” QFT RG terminology, where it exists only one operator $`X`$ of (7)–type (i.e., all $`A_j`$ except one are equal to zero), the procedure of group restriction on a particular BVP solution $`\overline{g}_{appr}`$ eliminates arbitrariness in the form of $`\beta (g)`$–function. While the general form of the condition given by Eq.(37) is the same for any BVP solution, the way of realization of the restriction procedure in every particular case employs a particular perturbation approximation (PA) for the concrete BVP. Generally, the restriction procedure reduces the dimension of $`𝒢`$. It also “fits” boundary conditions into the operator (35) by a special choice of coefficients $`A_j`$ and/or by choosing the particular form of arbitrary functions in coordinates $`\xi ^i`$, $`\eta ^\alpha `$. Hence, the general element (35) of the group $`𝒢`$ after the fulfillment of a restriction procedure is expressed as a linear combination of new generators $`R_i`$ with the coordinates $`\stackrel{~}{\xi }^i`$, $`\stackrel{~}{\eta }^\alpha `$, $$XR=\underset{j}{}B^jR_j,R_j=\stackrel{~}{\xi }_j^i_{x^i}+\stackrel{~}{\eta }_j^\alpha _{u^\alpha },$$ (38) where $`B^j`$ are arbitrary constants. The set of RGS generators $`R_i`$ each containing the desired BVP solution in its invariant manifold, define a group of transformations that we also refer to as renormgroup. Therefore, here we extend the notion of renormgroup and RG symmetry and the direct analogy with the “Bogoliubov RG” is preserved only for one-parameter group of point transformations. IV. The prescribed three steps entirely define the regular algorithm of RGS construction but do not touch on how a BVP solution is found. Hence, one more important, the fourth, step should be added. It consists in using RGS generators to find analytical expressions for the new, “improved”, solution of the BVP. Mathematically, this step makes use of RG=FS invariance conditions that are given by a combined system of (31) and the vanishing condition for the linear combination of coordinates $`\stackrel{~}{\kappa }_j^\alpha `$ of the canonical operator equivalent to (38), $$\underset{j}{}R^j\stackrel{~}{\kappa }_j^\alpha \underset{j}{}B^j\left(\stackrel{~}{\eta }_j^\alpha \stackrel{~}{\xi }_j^iu_i^\alpha \right)=0.$$ (39) One can see that conditions (39) are akin to (37). However, in contrast with the previous step, the differential variables $`u`$ in (39) should not be replaced by an approximate expression for the BVP solution $`U(z)`$, but should be treated as usual dependent variables. For the one-parameter Lie point renormgroup, RG invariance conditions lead to the first order PDE that gives rise to the so-called group invariants (like invariant couplings in QFT) which arise as solutions of associated characteristic equations. A general solution of the BVP is now expressed in terms of these invariants. On the one hand, this is in direct analogy with the structure of RG invariant solutions in QFT – compare with Eqs.(22) and (24). On the other hand, it reminds the so-called $`\mathrm{\Pi }`$–theorem from the theory of dimensional analysis and similitude (see, Section 19 in , Section 6 of Chapter 1 in and historical comment to Section 43 in ) directly related to power self-similarity, discussed above in Section 2.3. However, as we shall see later, in the general case of arbitrary RGS the group invariance condition obtained for BVP is not necessarily characteristic equations for the Lie point group operator. They may appear in a more complicated form, e.g., as a combination of PDE and higher order ODE (see Section 4.2). Nevertheless, the general idea of finding solution of the BVP as RG invariant solutions remains valid. ### 3.2 Examples of solution improving We present now a few examples of the RGS construction with further use of the symmetry for “improving” an approximate solution. #### 3.2.1 Modified Burgers equation As the first example, we take the initial value problem for the modified Burgers equation $$u_tau_x^2\nu u_{xx}=0,u(0,x)=f(x).$$ (40) It is connected with the heat equation $$\stackrel{~}{u}_t=\nu \stackrel{~}{u}_{xx}$$ (41) by transformation $`\stackrel{~}{u}=\mathrm{exp}(au/\nu )`$ and has an exact solution. Due to this, while using RGS to find a solution, one can check the validity of our approach. The RGS constructing for (40) is an apt illustration of the general scheme, shown in Figure 1 which may be helpful in understanding other examples of the general algorithm implementation. We review here in short the procedure and results of paper Ref.. The RG-manifold $``$ (step I) is given by Eq.(40) with the parameters of nonlinearity $`a`$ and dissipation $`\nu `$ included in the list of independent variables. The Lie calculational algorithm applied to $``$ gives, for the admitted group $`𝒢`$ (step II), nine independent terms in the general expression for the group generator $$X=\underset{i=1}{\overset{8}{}}A^i(a,\nu )X_i+\alpha (t,x,a,\nu )e^{au/\nu }_u,$$ (42) $$\begin{array}{c}X_1=4\nu t^2_t+4\nu tx_x(\nu /a)(x^2+2\nu t)_u,X_2=2t_t+x_x,\hfill \\ \hfill \\ X_3=(1/\nu )_t,X_4=2\nu t_x(\nu /a)x_u,X_5=_x,X_6=(\nu /a)_u,\hfill \\ \hfill \\ X_7=a_a+\left[(\nu /a)u\right]_u,X_8=2\nu _\nu +x_x+2\left[u(\nu /a)\right]_u.\hfill \end{array}$$ Here, $`A^i(a,\nu )`$ are arbitrary functions of their arguments and $`\alpha (t,x,a,\nu )`$ is an arbitrary function of four variables, satisfying the heat equation (41). A set of operators $`X_i`$ form an eight-dimensional Lie algebra $`L_8`$. The first six generators relate to the well-known symmetries of the modified (potential) Burgers equation (see, e.g., Vol.1, p.183 in ). They describe projective transformation in the $`(t,x)`$plane ($`X_1`$), dilatations in the same plane ($`X_2`$), translations along the $`t,x`$ and $`u`$axes ($`X_3`$, $`X_5`$ and $`X_6`$) and Galilean transformations ($`X_4`$). The last two generators $`X_7`$ and $`X_8`$ relate to dilatations of parameters $`a`$ and $`\nu `$ now involved in group transformations. The procedure of restriction (step III) of the group (42) admitted by $``$ (40) implies the check of the invariance condition (37) on a particular BVP solution $`u=U(t,x,a,\nu )`$ $$\left\{\eta _{\mathrm{}}+\underset{i=1}{\overset{8}{}}A^i(a,\nu )\kappa _i\right\}_{|u=U(t,x,a,\nu )}=0,\kappa _i\eta _i\xi _i^1u_t\xi _i^2u_x\xi _i^3u_a\xi _i^4u_\nu .$$ (43) This formula expresses the coordinate $`\alpha `$ of the last term in (42) via the remaining coordinates of eight generators $`X_i`$ for arbitrary $`t`$, and hence for $`t=0`$, when $`U(0,x,a,\nu )=f(x)`$. As a result, we obtain the “initial” value $`\alpha (0,x,a,\nu )`$ and then, using the standard representation for the solution to the linear parabolic equation (41), the value of $`\alpha `$ at arbitrary $`t0`$ $$\alpha (t,x,a,\nu )=\underset{i=1}{\overset{8}{}}A^i(a,\nu )<\overline{\kappa }_i(x,a,\nu )>.$$ (44) Here, $`\overline{\kappa }_i(x,a,\nu )`$ denote “partial” canonical coordinates $`\kappa _i`$ taken at $`t=0`$ and $`u=f(x)`$. Symbol $`<F>`$ designates the convolution of a function $`F`$ with the fundamental solution of (41), multiplied by the exponential function of $`f`$ entering into the boundary condition $$<F(x,t,a,\nu )>\frac{1}{\sqrt{4\pi \nu t}}\underset{\mathrm{}}{\overset{\mathrm{}}{}}𝑑yF(y,t,a,\nu )\mathrm{exp}\left(\frac{(xy)^2}{4\nu t}+\frac{af(y)}{\nu }\right).$$ Substitution (44) in the general expression (42) gives the desired RG generators $$R_i=X_i+\varrho _ie^{au/\nu }_u,$$ $$\begin{array}{c}\varrho _1=\frac{\nu }{a}<x^2>,\varrho _2=<xf_x>,\varrho _3=\frac{1}{\nu }<af_x^2+\nu f_{xx}>,\varrho _4=\frac{\nu }{a}<x>,\hfill \\ \hfill \\ \varrho _5=<f_x>,\varrho _6=\frac{\nu }{a}<1>,\varrho _7=<f\frac{\nu }{a}>,\varrho _8=<xf_x2f+2\frac{\nu }{a}>.\hfill \end{array}$$ Operators $`R_i`$ form an eight-dimensional RG algebra $`RL_8`$ that has the same tensor of structural constants as $`L_8`$, i.e. $`RL_8`$ and $`L_8`$ are isomorphic. Hence, the procedure of the group restriction eliminates the arbitrariness presented by the function $`\alpha `$ and “fits” the boundary conditions into RG generators by means of $`\varrho _i`$. It can be verified that the exact solution of the initial-value problem (40) $$u(t,x;a,\nu )=\frac{\nu }{a}\mathrm{ln}<1>\frac{\nu }{a}\mathrm{ln}\left\{\frac{1}{\sqrt{4\pi \nu t}}\underset{\mathrm{}}{\overset{\mathrm{}}{}}𝑑y\mathrm{exp}\left(\frac{(xy)^2}{4\nu t}+\frac{af(y)}{\nu }\right)\right\}$$ (45) is the invariant manifold for any of the above RGS operators. And vice versa, (45) can be reconstructed from an approximate solution with the help of any of the RGS operators or their linear combination. For example, two such operators, $`\nu R_3R_t`$ and $`(1/a)(R_6+R_7)R_a`$ were used in to reconstruct the exact solution from perturbative (in time and in nonlinearity parameter $`a`$) solutions. Below, we describe this procedure (step IV) using the operator $`R_a`$, $$R_a=_a+(1/a)\left(u+e^{au/\nu }<f(x)>\right)_u.$$ (46) It is evident that $`t,x`$ and $`\nu `$ are invariants of group transformations with (46), whilst finite RG transformations of the two remaining variables, $`a`$ and $`u`$, are obtained by solving the Lie equations for (46), with $`\mathrm{}`$ the group parameter $$\frac{da^{}}{d\mathrm{}}=1,a^{}|{}_{\mathrm{}=0}{}^{}=a;\frac{du^{}}{d\mathrm{}}=\alpha (t,x,a^{},\nu )e^{a^{}u^{}/\nu }\frac{u^{}}{a^{}},u^{}|{}_{\mathrm{}=0}{}^{}=u.$$ (47) Combining these equations yields one more invariant $`𝒥=e^{au/\nu }<1>`$ for the RGS generator (46). Solution of (47) along with (44) gives the formulae for finite RG transformations of the group variables $`\{t,x,a,\nu ,u\}`$ $$\begin{array}{c}t^{}=t,x^{}=x,\nu ^{}=\nu ,a^{}=a+\mathrm{},u^{}=\frac{\nu }{a+\mathrm{}}\mathrm{ln}\left(e^{au/\nu }+<e^{\mathrm{}f(x)/\nu }1>\right).\hfill \end{array}$$ (48) Choosing the value $`a`$ equal to zero, which is a starting point of PA in $`a`$, we get $`a^{}=\mathrm{}`$. Then after excluding $`t,x,\nu `$ and $`\mathrm{}`$ from the expression for $`u^{}`$ (48) and omitting accents over $`t^{},x^{},\nu ^{},u^{}`$ and $`a^{}`$ the desired BVP solution (45) is obtained. It also follows directly from $`𝒥`$ in view of the initial condition $`𝒥_{a=0}=0`$. A similar procedure can be fulfilled for the other RG operator, $$R_t=_t+e^{au/\nu }<af_x^2+\nu f_{xx}>_u,$$ which is consistent with the PA in time $`t`$. Although invariants for $`R_t`$ and finite RG transformations differ from that for (46), the final result, i.e., the exact solution of BVP (40) given by (45), is the same. This possibility is the distinct demonstration of the multi-dimensional RGS to reconstruct the unique BVP solution from different PA: either in parameter $`a`$ or in $`t`$ (though we used only two one–dimensional subalgebras here<sup>11</sup><sup>11</sup>11This can be considered as a construction parallel to the one used in Ref..). #### 3.2.2 BVP for ODE: simple example Quite recently, the QFT renormalization group ideology has been applied, a bit straightforward, in mathematical physics for asymptotic analysis of solutions to DEs and in constructing an envelope of the family of solutions . Our second methodological example with linear ODE is presented here in order to illustrate the difference between our approach and the “perturbative RG theory” devised in for a global analysis of BVP solutions in mathematical physics. Consider a linear second order ODE for $`y(t)`$ with the initial conditions at $`t=\tau `$, $$y_{tt}+y_t+\epsilon y=0,y(\tau )=\stackrel{~}{u},y_t(\tau )=\stackrel{~}{w},$$ (49) which has the exact solution: $$y=C_+e^{\gamma _+(t\tau )}+C_{}e^{\gamma _{}(t\tau )},\gamma _\pm =\frac{1\pm K}{2},K=\sqrt{14\epsilon },C_\pm =\frac{\stackrel{~}{w}+\gamma _{}\stackrel{~}{u}}{K}.$$ (50) Provided that the parameter $`\epsilon `$ is small, the solutions to Eq.(49) has been treated in with the aim to demonstrate effectiveness of the “perturbative RG theory” for an asymptotic analysis of a solution behaviour. The main goal of this treatment was to improve a perturbative expansion in powers of $`\epsilon `$ with secular terms $`\epsilon (t\tau )`$ and obtain<sup>12</sup><sup>12</sup>12 The algorithm used in for improving PA solutions with secular terms involves a) an introduction of some additional parameters in solutions, b) a special choice of these parameters that eliminates secular divergencies, and c) imposing a condition of independence of a solution upon the way of introducing these parameters. In some cases, this algorithm, directly borrowed from QFT RG-method, gives an exact solution. However, the question of correspondence of this construction to a transformation group of a solution of BEs remains open. a uniformly valid asymptotic of a solution $$y=c_+e^{(1+\epsilon (1+\epsilon ))(t\tau )}+c_{}e^{\epsilon (1+\epsilon )(t\tau )}+O(\epsilon ^2),$$ (51) $$c_+\left((1+2\epsilon )\stackrel{~}{w}+\epsilon \stackrel{~}{u}\right),c_{}\left((1+2\epsilon )\stackrel{~}{w}+(1+\epsilon )\stackrel{~}{u}\right),$$ which is accurate for small values $`\epsilon 1`$ but for arbitrary values of the product $`\epsilon (t\tau )`$. We are going to show that the use of our regular RG algorithm enables one to improve a PA solution (either in powers of $`\epsilon `$ or in $`t\tau `$) up to the exact BVP solution (50). Rewriting (49) in the form of the system of two first order ODEs for functions $`uy`$ and $`wy_t`$, $$u_t=w,w_t=\epsilon wu,$$ (52) we construct $``$ (step I) using the invariant embedding method (this approach has first been realized in ). Then, $``$ is presented as a joint system of BEs (52) and embedding equations $$u_\tau (\epsilon \stackrel{~}{w}+\stackrel{~}{u})u_{\stackrel{~}{w}}\stackrel{~}{w}u_{\stackrel{~}{u}}=0,w_\tau (\epsilon \stackrel{~}{w}+\stackrel{~}{u})w_{\stackrel{~}{w}}\stackrel{~}{w}w_{\stackrel{~}{u}}=0,$$ treated in the extended space of group variables which include the parameters $`\tau ,\stackrel{~}{w},\stackrel{~}{u}`$ of boundary conditions in addition to $`t`$ and dependent variables $`u,w`$. Omitting tedious calculations related to the following two steps (steps II and III ), we present here two examples of resulting RGS generators $$\begin{array}{c}R_\tau =_\tau (\stackrel{~}{w}+\epsilon \stackrel{~}{u})_{\stackrel{~}{w}}+\stackrel{~}{w}_{\stackrel{~}{u}},\hfill \\ \hfill \\ R_\epsilon =_\epsilon \left(\frac{1}{\mu ^2}(2w+u)(1t/2)+(t/2)u\right)_w+\frac{t}{\mu ^2}\left(2w+u\right)_u\hfill \\ \hfill \\ \left(\frac{1}{\mu ^2}(2\stackrel{~}{w}+\stackrel{~}{u})(1\tau /2)+(\tau /2)\stackrel{~}{u}\right)_{\stackrel{~}{w}}+\frac{\tau }{\mu ^2}\left(2\stackrel{~}{w}+\stackrel{~}{u}\right)_{\stackrel{~}{u}},\hfill \end{array}$$ (53) that involve the initial values $`\stackrel{~}{w}`$, $`\stackrel{~}{u}`$ and initial point $`\tau `$ in RG transformations. In addition, $`R_\epsilon `$ transforms the parameter $`\epsilon `$. Now, the procedure of constructing the BVP solution (50) (step IV) is similar to that used in the previous Section 3.2.1 and employ finite transformations that are defined by the Lie equations for the operators (53). For $`R_\tau `$ functions $`u,w`$ and the parameter $`\epsilon `$ are group invariants, while the translations of $`\tau `$ and the corresponding transformations of $`\stackrel{~}{u},\stackrel{~}{w}`$ restores the exact solution (50) from the PA in powers of $`t\tau `$ (note that the parameter $`\epsilon `$ is not necessarily small in this PA!). For $`R_\epsilon `$ the difference $`t\tau `$ is group invariant, whilst the transformation of $`\epsilon `$ and related transformations of $`u,w,\stackrel{~}{u},\stackrel{~}{w}`$ restore the exact solution (50) from the PA (discussed in ) in powers of $`\epsilon `$. Hence, as in the previous Section 3.2.1, both the RGS generators (53) reconstruct the unique BVP solution from different PAs. ## 4 RG in Nonlinear optics ### 4.1 Formulation of a problem As a problem of real physical interest, take BVP that describes self-focusing of a high-power light beam. While the problem plays an important role in nonlinear electrodynamics since 60s, the detailed quantitative understanding of self-focusing is still missing , and there is no method which allows to find an analytic solution to the corresponding equations with arbitrary boundary conditions. Here, we demonstrate the great potential of the RGS approach in constructing analytic solutions of BVP equations with arbitrary boundary conditions. The RGS method allows to consider different types of BEs for self-focusing process which include plane and cylindrical beam geometry, nonlinear refraction and diffraction. The merit of the RGS method is that it describes BVP solutions with one– or two–dimensional singularities in the entire range of variables from the boundary up to the singularity point. Let us start with BVP for the system of two DEs $$v_z+vv_x\alpha n_x=0,n_z+nv_x+vn_x+(\nu 1)\left(nv/x\right)=0,$$ (54) $$v(0,x)=0,n(0,x)=N(x),$$ (55) which are used in nonlinear optics of self-focusing wave beams when diffraction is negligible. We study spatial evolution of the derivative of the beam eikonal $`v`$ and the beam intensity $`n`$ in the direction inwards the medium $`z`$ and in the transverse direction $`x`$. The term proportional to $`\alpha `$ is related to nonlinear refraction effects; $`\nu =1`$ and $`\nu =2`$ refers to the plane and cylindrical beam geometry, respectively. Boundary conditions (55) correspond to the plane front of the beam and the arbitrary transverse intensity distribution. ### 4.2 Plane geometry In the plane beam geometry (at $`\nu =1`$) Eqs.(54) can be reduced to the system of BEs $$\tau _wn\chi _n=0,\chi _w+\alpha \tau _n=0,$$ (56) for functions $`\tau =nz`$ and $`\chi =xvz`$ of $`w=v/\alpha `$ and $`n`$ arguments, with boundary conditions $$\tau (0,n)=0,\chi (0,n)=H(n),$$ (57) where $`H(n)`$ is the inverse to $`N(x)`$. Here, the procedure of RGS constructing makes use of the Lie–Bäcklund symmetry and is described as follows . The manifold $``$ (step I) is defined by Eqs.(56) treated in the extended space that include dependent and independent variables $`\tau ,\chi ,w,n`$ and derivatives of $`\tau `$ and $`\chi `$ with respect to $`n`$ of an arbitrary high order. The admitted symmetry group $`𝒢`$ (step II) is represented by the canonical Lie-Bäcklund operator $$X=f_\tau +g_\chi ,$$ (58) with the coordinates $`f`$ and $`g`$ that are linear combinations of $`\tau `$ and $`\chi `$ and their derivatives $`^i\tau /n^i`$ and $`^i\chi /n^i,i1`$ with the coefficients depending on $`w`$ and $`n`$. The restriction of the group admitted by $``$ (56) (step III)) implies the check of the invariance condition (37) that yields two relations $$f=0,g=0.$$ (59) These relations should be valid on a particular solution of BVP with the boundary data (57). For example, choosing the so-called “soliton” profile, $`N(x)=\mathrm{cosh}^2(x)`$, i.e., $`H(n)=\mathrm{Arccosh}\left(1/\sqrt{n}\right)`$, we have $$\begin{array}{cc}f=\hfill & 2n(1n)\tau _{nn}n\tau _n2nw(\chi _n+n\chi _{nn})+\left(\alpha w^2/\mathrm{\hspace{0.17em}2}\right)n\tau _{nn},\hfill \\ \hfill & \\ g=\hfill & 2n(1n)\chi _{nn}+(23n)\chi _n+\alpha w\left(2n\tau _{nn}+\tau _n\right)+\left(\alpha w^2/\mathrm{\hspace{0.17em}2}\right)\left(n\chi _{nn}+\chi _n\right).\hfill \end{array}$$ (60) Dependence on $`\tau _{nn}`$ and $`\chi _{nn}`$ indicates that here RGS is the second-order Lie-Bäcklund symmetry. In order to find a particular solution of a BVP (step IV), one should solve the joint system of BEs (56) and second-order ODEs that follow from the RG=FS invariance conditions (59) and (60). The resulting expressions – the well-known Khokhlov solutions<sup>13</sup><sup>13</sup>13In Ref., where this solution was first obtained, it did not result from a regular procedure. $$v=2\alpha nz\mathrm{tanh}(xvz),\alpha n^2z^2=n\mathrm{cosh}^2(xvz)1,$$ (61) describe the process of self-focusing of a soliton beam: the sharpening of the beam intensity profile with the increase of $`z`$ is accompanied by the intensity growth on the beam axis. The solution (61) is valid up to the singularity point where the derivatives $`v_x`$ and $`n_x`$ tend to infinity whilst the beam intensity $`n`$ remains finite $$z_{sing}^{sol}=1/2\sqrt{\alpha },n_{sing}^{sol}=2.$$ (62) Here, the Lie-Bäcklund RGS enables one to reconstruct the BVP solution and describe the solution singularity for the light beam with the soliton initial intensity profile. One more example of an exact BVP solution obtained with the help of Lie-Bäcklund RGS (with the initial beam profile in the form of a ”smoothed” step) can be found in . For arbitrary boundary data, it turns to be impossible to fulfill the condition (59) with the help of the Lie-Bäcklund symmetries of any finite order, and one is forced to use a different algorithm of RGS constructing, based on the approximate group methods. Here, (step I)) $``$ is given by BEs (56) with a small parameter $`\alpha `$, and coordinates of the group generator (58) (and, hence, coordinates of the RGS operators) appear as infinite series in powers of $`\alpha `$ $$f=\underset{i=0}{\overset{\mathrm{}}{}}\alpha ^if^i;g=\underset{i=0}{\overset{\mathrm{}}{}}\alpha ^ig^i.$$ (63) The procedure of finding the coefficients $`f^i,g^i`$ (step II)) leads to the system of recurrent relations that express higher-order coefficients $`f^{i+1}`$, $`g^{i+1}`$ in terms of previous ones $`f^i`$, $`g^i`$. It means that once the zero-order terms are specified, the other terms are reconstructed by the recurrent relations. The coefficients $`f^i`$ and $`g^i`$ contain an arbitrary function of $`n,\chi _{[s]}`$ and $`\tau _{[s]}w(s\chi _{[s]}+n\chi _{[s+1]})`$ where subscript $`[s]`$ denotes the partial derivative of the order $`s`$ with respect to $`n`$. This arbitrariness is eliminated by the procedure of group restriction (step III)), i.e., by imposing the invariance condition (59). For particular forms of $`f^0`$ and $`g^0`$, that is for partial boundary conditions (57), infinite series are truncated automatically, and we arrive at the exact RGS. One example of this kind is given by Eqs.(60) that have a binomial structure $`f=f^0+\alpha f^1`$, $`g=g^0+\alpha g^1`$. If we neglect the higher-order terms in the case of arbitrary boundary conditions (when series (63) are not truncated automatically), then we get an approximate RGS which produces an approximate solution to the BVP. As an example, we give here two sets of expressions for the coordinates $`f^i`$ and $`g^i`$ for the Gaussian initial profile with $`N(x)=\mathrm{exp}(x^2)`$, i.e., $`H(n)=\left(\mathrm{ln}(1/n)\right)^{1/2}`$, which define approximate RGS $$\begin{array}{cc}𝐚)\hfill & f^0=1+2n\chi \chi _n,g^0=0,f^1=2\tau \tau _n+\frac{\tau ^2}{n},g^1=2\left(\tau \chi _n+\chi \tau _n\right),\hfill \\ \hfill & \\ 𝐛)\hfill & f^0=2n(\tau \chi _n+\tau _n\chi ),g^0=1+2n\chi \chi _n,f^1=2\chi \tau _\alpha ,g^1=2\left(\chi \chi _\alpha \tau \tau _n\right).\hfill \end{array}$$ Here, linear dependencies of $`f`$ and $`g`$ upon first-order derivatives indicate that RGS is equivalent to Lie point symmetry. The peculiarity of the case $`𝐛`$ is a dependence of $`f`$ and $`g`$ not only on derivatives with respect to $`n`$ but also with respect to $`\alpha `$: it means that the parameter $`\alpha `$ is also involved in group transformations. In the non-canonical representation (33), the RGS generator in this case has the form $$R_{Guass1}=2\tau _w+2n\chi _n+2\alpha \chi _\alpha _\chi .$$ (64) The last step IV is performed in a usual way by solving the joint system of BEs (56) and equations that follow from the RG=FS invariance condition (59), or else, using invariants of associated characteristic equations for RG operator provided that RGS is a Lie point symmetry. We give here the solution that follows from RGS (64), $$x^2=\left(12\alpha nz^2\right)^2\mathrm{ln}\frac{1}{n(1\alpha nz^2)},v=\frac{2x\alpha nz}{12\alpha nz^2}.$$ (65) These expressions describe a self-focusing Gaussian beam (the plot $`n(x)`$ for this solution is presented at the end of the section on Figure 2), that is qualitatively very similar to the spatial evolution of the soliton beam (61). Moreover, the singularity position and the value of maximum beam intensity at this point coincide with analogous values (62) for the soliton beam. Although formulae (65) correspond to an approximate BVP solution, they exactly describe the behaviour of $`n`$ on the beam axis at $`x=0`$. To estimate the reliability of result (65) in the off-axis region, we compared it with another approximate BVP solution which arises from the approximate RGS in the case $`𝐚`$. These approximations agree very well (details are presented in ), thus proving the accuracy<sup>14</sup><sup>14</sup>14One more evidence is provided by the comparison of approximate and exact BVP solution for the soliton beam performed in . of the RG approach. ### 4.3 Cylindrical geometry In the above discussion we dealt with the plane beam geometry and took into account only effects of nonlinear beam refraction, neglecting diffraction. The flexibility of RGS algorithm allows one to apply it in a similar way to a more complicated model as compared to (56), e.g., for the cylindrical beam geometry, $`\nu =2`$. Omitting technical details, we present the RGS generator for the cylindrical parabolic beam with $`N=1x^2`$ $$R_{par}=\left(12\alpha z^2\right)_z2\alpha zx_x2\alpha \left(xvz\right)_v+4\alpha nz_n.$$ The BVP solution is expressed in terms of group invariants for this generator: $$𝒥_1=\frac{x^2}{\varrho };𝒥_2=n\varrho ;𝒥_3=2\alpha x^2v^2\varrho +\frac{xv}{2}\varrho _z;\varrho =\left(12\alpha z^2\right).$$ (66) The explicit form of dependencies of $`𝒥_2=1𝒥_1^2,𝒥_3=2\alpha 𝒥_1`$ upon $`𝒥_1`$ follows from the boundary conditions (55). They lead to the well-known solution $$v=(x/(2\varrho ))\varrho _z,n=(1/\varrho )(1(x^2/\varrho ))$$ (67) that describes the convergence of the beam to the singularity point $`z_{sing}^{par}=1/\sqrt{2\alpha }`$ where $`\varrho =0`$ and $`n\mathrm{}`$. The solution singularity is two-dimensional here: the infinite growth of beam intensity in the vicinity of the singularity $`zz_{sing}^{par}`$ is accompanied by the infinite growth of the derivative $`v_x`$ and collapsing of the beam size in the transverse direction. The RGS algorithm based on approximate group methods can also be applied in the case when besides nonlinear refraction also diffraction effects are taken into account. Then, the first equation in (54) should be modified by adding the diffraction term $$\beta _x\left\{\left(x^{1\nu }/\sqrt{n}\right)_x\left(x^{\nu 1}_x\sqrt{n}\right)\right\}.$$ Standard calculations done in compliance with a general scheme for thus modified $``$ (for details see ) give the RGS generator for the cylindrical beam geometry $`(\nu =2)`$ $$R_{Gauss2}=\left(1+z^2S_{\chi \chi }\right)_z+\left(zS_\chi +vz^2S_{\chi \chi }\right)_x+S_\chi _v\left[nz\left(1+\frac{vz}{x}\right)S_{\chi \chi }+\frac{nz}{x}S_\chi \right]_n.$$ (68) Here the function $`S`$, defined by the form of the intensity boundary distribution, $$S(\chi )=\alpha N(\chi )+\frac{\beta }{\chi \sqrt{N(\chi )}}_\chi \left(\chi _\chi \sqrt{N(\chi )}\right),$$ contains two small parameters, $`\alpha `$ and $`\beta `$, and, as in the case $`\beta =0`$, there exist specific forms of boundary distribution, $`N`$, for which the RGS operator (68) defines exact (not approximate) symmetry valid for arbitrary values of $`\alpha `$ and $`\beta `$. Constructing a particular BVP solution (step IV) implies the use of group invariants related to (68), and the procedure is similar to that one for the parabolic beam. For the Gaussian wave beam, $`N=\mathrm{exp}(x^2)`$, the result is as follows: $$v(z,x)=\frac{x\chi }{z},n(z,x)=e^{\mu ^2}\frac{\chi }{x}\frac{\left(\beta \alpha e^{\chi ^2}\right)}{\left(\beta \alpha e^{\mu ^2}\right)}.$$ (69) Here $`\chi `$ and $`\mu `$ are expressed in terms of $`t`$ and $`x`$ by the implicit relations $$\begin{array}{c}\beta \left(\mu ^2\chi ^2\right)+\alpha \left(e^{\mu ^2}e^{\chi ^2}\right)=2z^2\chi ^2\left(\beta \alpha e^{\chi ^2}\right)^2;\hfill \\ \hfill \\ x=\chi \left(1+2z^2\left(\beta \alpha e^{\chi ^2}\right)\right).\hfill \end{array}$$ The solution (69) describes the self-focusing of the cylindrical Gaussian beam that gives rise to the two-dimensional singularity: both the beam intensity $`n`$ and derivatives $`v_x,n_x`$ go to infinity at the point $`z_{sing}^{Gauss}=1/\sqrt{2(\alpha \beta )}`$ provided that $`\alpha >\beta `$. A detailed analysis of (69) and more general solutions with a parabolic form of an eikonal at $`z=0`$, $`v(0,x)=x/T`$, is given in . To illustrate the difference between the one– and two–dimensional solution singularities, in Figure 2 we present a typical behavior of the wave beam intensity, defined by Eq.(65) and (69). The left panel corresponds to the plane beam geometry, $`\nu =1`$, and without diffraction, $`\beta =0`$, while the right one is concerned with a cylindrical wave beam, $`\nu =2`$, with both nonlinearity and diffraction effects included. Diverse curves describe beam intensity distribution upon coordinate $`x`$ at different distances from the medium boundary, where we have the collimated Gaussian beam, $`N=\mathrm{exp}(x^2)`$. It is clear that in the plane geometry the derivative of the beam intensity with respect to $`x`$ turns to infinity at some singular point, while the value of intensity on the axis remains finite. In cylindrical case the solution singularity is two-dimensional: both the beam intensity and its derivative with respect to transverse coordinate turn to infinity simultaneously at point $`z_{sing}`$. This last example demonstrates the possibility of the RGS approach to analyse two-dimensional singularity. In the practice of RG application to critical phenomena, this correlates with the case of “two renormalization groups”. ## 5 Overview To complete our review, we indicate milestones in evolution of the RG concept. Since its appearance in QFT, RG served as a powerful tool of analyzing diverse physical problems and improving solution singularities disturbed by perturbation approximation. The development of the RG concept can be divided into two stages. The first one (since the mid-50s up to the mid-80s) is summarized in Figure 3. Besides early history (discovery of RG, formulation of the RG method and application to UV and IR asymptotics), it comprises the devising of the Kadanoff–Wilson RG in the 70s and following explosive expansion into other fields of theoretical physics. During this stage, the formulation of the RG method was based on the unified scaling transformation of an independent variable (and/or some parameters) accompanied by a more complicated transformation of a solution characteristic $`g_\mu =\overline{g}(\mu ,g)`$ – see Eqs.(10), (13) and (16) in Section 1.2. Here, the main role of RG=FS was the á priori establishment of the fact that the solution under consideration admits functional transformations that form a group. Any particular implementation of the RG symmetry differs in the form of the function(s) $`\overline{g}(\mu ,g)`$ (or $`\beta (g)`$) which, in an every partial case, is obtained from some approximate solution. The next stage, after the mid–80s, is depicted in the Figure 4. The scheme describes the entire evolution of the Bogoliubov RG. There were several important reasons in further devising the RG concept in theoretical physics at this period. On the one hand, it was due to the extension of the notion of FS and RG symmetry that until then were based on one-parameter Lie group of point transformations. Appending multi-dimensional Lie point groups and Lie–Bäcklund groups to possible realization of group symmetry enhanced the capability of RG method. On the other hand, this additional possibility arose due to the mathematical apparatus that was used in mathematical physics to reveal RGS. The advantage came from infinitesimal transformations that enabled to describe RGS by an algebra of RG generators. However, in contrast to the situation typical of QFT models with only one operator, in mathematical physics we have finite or infinite-dimensional algebras. Both their dimension and the method of construction depend upon a model employed and upon a form of boundary conditions. The use of infinitesimal approach results in constructing the RG-type symmetry with the help of regular methods of group analysis of DEs. Precisely, this regular algorithm naturally includes the RG=FS invariance condition in the general scheme of constructing and application of RGS generators (see also our recent review ). Within the infinitesimal approach this condition is formulated in terms of vanishing of canonical RG operator coordinates, which is especially important for Lie–Bäcklund RGS because finite transformations in this case are expressed as formal series. In particular, this property attribute a new feature to the RG analysis of a BVP solution with singular behavior, making a singularity analysis more powerful. At the same time, as the group analysis technique is still developing – here we mean both extension to new types of symmetries and application to more complicated mathematical models, e.g., including integro–differential equations – we have a clear perspective that the possibilities of a regular scheme based upon the Bogoliubov renormalization group method are far from being exhausted. Acknowledgments The authors are grateful to Professors Cris Stephens and Denjoe O’Connor for invitation to participate in the Conference “Renormalization Group 2000”. They are indebted to these gentlemen for useful discussions and comments. This work was partially supported by grants of the Russian Foundation for Basic Research (RFBR projects Nos 96-15-96030, 99-01-00232 and 99-01-00091) and by INTAS grant No 96-0842, as well as by the Organizing Committee of the above-mentioned meeting.
warning/0001/astro-ph0001151.html
ar5iv
text
# 1 Introduction ## 1 Introduction The Lake Baikal Neutrino Experiment exploits the deep water of the great Siberian lake as a detection medium for high energy neutrinos via muons and electrons, generated in high energy neutrino interactions. High energy neutrinos are of particular importance for high energy astrophysics for the last few decades to shed light on the physics of Active Galactic Nuclei (AGN), binary star systems, gamma ray burst (GRB) etc. The life time of the lake Baikal Neutrino Experiment spans almost two decades, from first small experiment with a few PMTs to the present large scale neutrino telescope NT-200 <sup>?,?,?)</sup>, which has been put into full operation on April 6th 1998. The effective area of the telescope for muons is 2000-10000 m<sup>2</sup> depending on the muon energy. The expected rate of muons from atmospheric neutrinos, with a muon energy thresholds of 10 GeV and after all cuts rejecting background, is about 1 per two days. ## 2 Neutrino Telescope NT-200 ### 2.1 Detector and Site The Neutrino Telescope NT-200 is located in the southern part of Lake Baikal (51.50N and 104.20E) at 3,6km from the shore and at a depth of 1km. The absorption length $`L_{abs}`$ of water at the site for wavelengths between 470nm and 500nm is about 20m and seasonal variations are less than 20%. The light scattering is subjected strongly to seasonal variations and from year to year. We can say now that light scattering is rather strongly anisotropic and typical values of $`L_{scatt}`$ are about 15m. NT-200 (fig.1) consists of 192 optical modules (OMs) at 8 strings arranged at an umbrella like frame <sup>?,?)</sup>. Pairs of OMs are switched in coincidence with 15ns time window and define a channel. We pursue pairwise ideology for many reasons: to suppress individual OM background counting rates due to OM dark current and water luminescence, the level of late- and afterpulses etc. The OMs <sup>?)</sup> consists of QUASAR-370 phototube <sup>?,?)</sup> enclosed in transparent, nearly spherical pressure housing. The optical contact between the photocathode region of the phototube and the pressure sphere is made by liquid glycerine sealed with a layer of polyurethane. Apart from the phototube every OM contains two HV power supplies (25 kV and 2 kV), a voltage divider, two preamplifiers, a calibration LED and a vacuum probe. The QUASAR-370 phototube has been developed especially for the Lake Baikal Neutrino Experiment by INR and KATOD Company in Novosibirsk. The phototube is a hybrid one and has excellent time and amplitude resolutions. The detector electronics system <sup>?)</sup> is hierarchical: from the lowest level to the highest one – OM’s electronics, ”sviazka” electronics module, string electronics module and the detector electronics module, where detector trigger signals are formed and all information from the string electronics module are received and sent to the shore station. The detector is operated from the shore station. A muon trigger is formed by the requirement of $`N`$ hits (with hit referring to a channel) within 500 ns. $`N`$ is typically set to the value 3 or 4. For such events, amplitude and time of all fired channels are digitized and sent to the shore. The event records includes all hits within a time window of -1000 ns to +800 ns with respect to the muon trigger signal. A separate monopole trigger system searches for clusters of sequential hits in individual channels which are characteristic for the passage of slowly moving, bright objects like GUT monopoles. There is a separate hydrological string to study permanently water parameters of the lake. This string is deployed at about 60m from the main part of NT-200. There are two nitrogen lasers for the detector time calibration placed just above and below the detector. The former one illuminates each individual channel via fiber optics and the latter illuminates the detector as a whole. ### 2.2 Cherenkov EAS Array To determine angular resolution of the NT-200 mobile wide angle air cherenkov array has been developed <sup>?)</sup>. This array (Fig.2) has been deployed for the last two expeditions on the ice just above NT-200. It consists of four optical modules put on the sledges for fast deployment. Three of them are fixed in the vertices of an equilateral triangle and another one just in the center of the triangle. The sidelength of the triangle is about 100 m. The analog signals from the optical modules are fed by electrical coax cables to the central electronic station which is located near the central optical module. The data acquisition system includes 4 constant fraction discriminators, a majority coincidence unit, two TDCs with 500 ns and 5000 ns ranges, ADC, an EAS event counter, counting rate scalers for each channel and an underwater master signals counter. Each optical module incorporates QUASAR-370G phototubes, two high voltage power supplies of 25 kV and 1 kV for an electron-optical preamplifier and small PMT respectively, anode pulses preamplifier and LED for amplitude calibration. The phototube is arranged into light tight metallic box which is equipped with mechanically removable lid. To increase the sensitivity area, Winston cones are used providing almost 2450 cm<sup>2</sup> final sensitive area by a factor of 2 larger than in the 1998 array. The angular acceptance of the optical module is restricted to 30<sup>0</sup> of half angle. The QUASAR-370G phototube is practically the same as the phototube used in NT-200 but with six stages small PMT developed to withstand high mean anode current due to night sky background (NSB). To stabilize the phototubes gain, the HV power supplies are surrounded by a thermostat. A 4-fold coincidence within a 1000 ns gate defines the trigger of the array. Using light concentrators allows to increase the array trigger rate by a factor of two. The trigger rate depends on the weather conditions and is in average to 0.8-1Hz. The trigger signal of underwater telescope is fed via more than 1km long coax cable to the center electronic station of the Cherenkov array and switched into coincidence with the array trigger signal. The time difference between them is measured by a wide range TDC. The synchronization between Cherenkov array and underwater telescope is done by comparing two underwater event counters in the central electronic station on the ice and detector electronic module of the underwater telescope read out from the shore station. The array energy threshold is about 100 TeV. The angular resolution of the array is $`0.5^{}÷1^{}`$ giving a good reference point to estimate the ngular resolution of the underwater telescope for muons close to the vertical direction ($`0^{}÷30^{}`$) since high energy muons retain the direction of their parent shower. In 1998 the EAS array operated in coincidence with only one string, and overall 450 events were registered. Analysis of those events showed that the accuracy of zenith angle reconstruction taking into account only time information is close to $`6^{}`$, what is in reasonable agreement with MC calculations ($`5^{}`$). In 1999 we managed to establish the joint work of EAS array and NT-200, but unfortunately we collected only 150 events due to bad weather. The analysis of data is still in progress and results will be presented later. ### 2.3 Shadow of the shore in muons Muon angular distributions as well as depth dependence of the vertical flux obtained from data taken with NT-36 have been presented earlier <sup>?)</sup>. Another example which confirms the efficiency of track reconstruction uses the shore ”shadow” in muons recorded with NT-96. As it was already mentioned, NT-200 is situated at a distance of 3.6 km to the nearby shore of the lake, and more than 30 km to the opposite shore. This asymmetry allows to study the asymmetry in the azimuth distribution of muons under large zenith angles, where reconstruction for the rather ”thin” NT-96 is most critical. A sharp decrease of the muon intensity at zenith angles of $`70^{}÷90^{}`$ is expected. The comparison of the experimental muon angular distribution with MC calculations gives us an estimation of the accuracy of the reconstruction error close to the horizontal direction. Indeed, the NT-96 data show clear dip of the muon flux in the direction of the shore and for zenith angles larger than $`70^{}`$ (fig.3) – in very good agreement with calculations which take into consideration the effect of the ”shadowing” shore. ## 3 Selected physics results In this chapter we present some physics results based on the data of 70 days effective operating time of the 4-string array NT-96. ### 3.1 Separation of neutrino events with full track reconstruction The signature of neutrino induced events is a muon crossing the detector from below. With the flux of downward muons exceeding that of upward muons from atmospheric neutrino interactions by about 6 orders of magnitude, a careful reconstruction is of prime importance. The reconstruction algorithm <sup>?)</sup> is based the assumption that the light induced by muons is emitted exactly under the Cherenkov angle ($`42^{}`$) relative to the muon track. For full track reconstruction ($`\theta `$, $`\varphi `$ and spatial coordinates) one needs more than 5 hit channels on more than 3 strings. In contrast to first stages of the detector (NT-36 <sup>?)</sup>), NT-96 can be considered as a real neutrino telescope for a wide region in zenith angle $`\theta `$. After the reconstruction of all events with $``$ 9 hits at $``$ 3 strings (trigger 9/3), quality cuts have been applied in order to reject fake events. Furthermore, in order to guarantee a minimum lever arm for track fitting, events with a projection of the most distant channels on the track ($`Z_{dist}`$) less than 35 meters have been rejected. Due to the small transversal dimensions of NT-96, this cut excludes zenith angles close to the horizon. The efficiency of the procedure has been tested with a sample of $`1.8\times 10^6`$ MC-generated atmospheric muons, and with MC-generated upward muons due to atmospheric neutrinos. It turns out that the signal to noise ratio is $`>1`$ for this sample. The reconstructed angular distribution of $`2\times 10^7`$ events taken with NT-96 in April/September 1996 – after all cuts – is shown in Fig.4. From 70 days of NT-96 data, 12 neutrino candidates have been found. Nine of them have been fully reconstructed. Three nearly upward vertical tracks (see subsection 3.2) hit only 2 strings and give a clear zenith angle but ambiguities in the azimuth angle – similar to the two events from NT-36 <sup>?)</sup>. This is in good agreement with MC expectations. ### 3.2 Search for nearly vertical upward moving neutrinos Unlike the standard analysis <sup>?)</sup>, the method presented in this section relies on the application of a series of cuts which are tailored to the response of the telescope to nearly vertically upward moving muons <sup>?,?)</sup>. The cuts remove muon events far away from the opposite zenith as well as background events which are mostly due to pair and bremsstrahlung showers below the array and to naked downward moving atmospheric muons with zenith angles close to the horizon ($`\theta >60^{}`$). The candidates identified by the cuts are afterwards fitted in order to determine the zenith angle. We included all events with $``$4 hits along at least one of all hit strings. To this sample, a series of 6 cuts is applied. Firstly, the time differences of hit channels along each individual string have to be compatible with a particle close to the opposite zenith(1). The event length should be large enough(2), the maximum recorded amplitude should not exceed a certain value(3), and the center of gravity of hit channels should not be close to the detector bottom(4). The latter two cuts reject efficiently brems showers from downward muons. Finally, also time differences of hits along different strings have to correspond to a nearly vertical muon (5) and the time difference between top and bottom hit in an event has to be larger than a minimum value (6). The effective area for muons moving close to opposite zenith and fulfilling all cuts exceeds $`1000`$ m<sup>2</sup>. Within 70 days of effective data taking, $`8.4\times 10^7`$ events with the muon trigger $`N_{hit}4`$ have been selected. Table 1 summarizes the number of events from all 3 event samples (MC signal and background, and experiment) which survive the subsequent cuts. After applying all cuts, four events were selected as neutrino candidates, compared to 3.5 expected from MC. One of the four events has 19 hit channels on four strings and was selected as neutrino candidate by the standard analysis too. The zenith angular distribution of these four neutrino candidates is shown in the inner box of Fig.3. Regarding the four detected events as being due to atmospheric neutrinos, one can derive an upper limit on the flux of muons from the center of the Earth due to annihilation of neutralinos - the favored candidate for cold dark matter. The limits on the excess muon flux obtained with underground experiments <sup>?,?,?)</sup> and NT-96 are shown in Table 2. The limits obtained with NT-96 are 4–7 times worse then the best underground limits since only the first 70 days of NT-96 have been analysed. This result, however, illustrates the capability of underwater experiments with respect to the search for muons due to neutralino annihilation in the center of the Earth. MC shows that for NT-200 the effective area is about 2000 m<sup>2</sup>, for $`E_\mu `$$`>`$10GeV, two times larger than for NT-96. In case that the energy threshold for upward muons could be decreased to 5 GeV NT-200 will permit to select a non-negligible amount of contained events and estimate the energy of muons. This will allow to study the neutrino energy spectrum for neutrinos having crossed about 13000 km in the Earth <sup>?)</sup>. Estimates show that the full number of contained events will be about 20 for $`\theta >165^{}`$ per year. In case of $`\nu _\mu `$$`\nu _x`$ oscillations, the $`\nu _\mu `$ flux will be suppressed and for $`\mathrm{\Delta }m^2`$=$`10^3`$eV<sup>2</sup> we will find only 7 events. ### 3.3 High Energy Neutrinos The ultimate goal of large underwater neutrino telescopes is the identification of extraterrestrial high energy neutrinos. In this chapter we present results of a search for neutrinos with $`E_\nu >10`$TeV derived from NT-96 data. Cherenkov light emitted by the electro-magnetic and (or) hadronic particle cascades and high energy muons produced at the neutrino interaction vertex in a large volume around the neutrino telescope. Earlier, a similar strategy has been used by the DUMAND <sup>?)</sup> and the AMANDA <sup>?)</sup> collaborations to obtain upper limits on the diffuse flux of high energy neutrinos. Fig.5 illustrates the detection principle. We select events with high multiplicity of hit channels corresponding to bright cascades. The volume considered for generation of cascades is essentially below the geometrical volume of NT-96 – its upper plane crosses the center of the telescope. A cut is applied which accepts only time patterns corresponding to upward traveling light signals (see below). This cut rejects most events from brems-cascades along downward muons since the majority of muons is close to the vertical; they would cross the detector and generate a downward time pattern. Only the fewer muons with large zenith angles may escape detection and illuminate the array exclusively via bright cascades below the detector. These events then have to be rejected by a stringent multiplicity cut. Neutrinos produce showers and high energy muons through CC-interactions $$\nu _l(\overline{\nu _l})+N\stackrel{CC}{}l^{}(l^+)+\text{hadrons},$$ (1) through NC-interactions $$\nu _l(\overline{\nu _l})+N\stackrel{NC}{}\nu _l(\overline{\nu _l})+\text{hadrons},$$ (2) where $`l=e`$ or $`\mu `$, and through resonance production <sup>?,?,?)</sup> $$\overline{\nu _e}+e^{}W^{}\text{anything},$$ (3) with the resonant neutrino energy $`E_0=M_w^2/2m_e=6.3\times 10^6`$GeV and cross section $`5.02\times 10^{31}`$cm<sup>2</sup>. Within the first 70 days of effective data taking, $`8.4\times 10^7`$ events with the muon trigger $`N_{hit}4`$ have been selected. For this analysis we used events with $``$4 hits along at least one of all hit strings. The time difference between any two channels deployed on the same string was required to obey the condition: $$(t_it_j)z_{ij}/c<az_{ij}+2\delta ,(i<j).$$ (4) The $`t_i,t_j`$ are the arrival times at channels $`i,j`$, and $`z_{ij}`$ is their vertical distance. With $`\delta =5`$nsec accounting for the timing error, the condition $`(t_it_j)z_{ij}/c<2\delta `$ (i.e. $`a=0`$) would cut for a signal traveling vertically upward with the speed of light, $`c`$. Setting $`a`$ to $`1`$ nsec/m, the acceptance cone around the opposite zenith is slightly increased. This condition has been used for almost vertically up-going muons selection earlier <sup>?,?)</sup>. 8608 events survive the selection criterion (4). Fig.6 shows the hit multiplicity distribution for these events (dashed histogram) as well as the expected one for background showers produced by atmospheric muons (solid histogram). The experimental distribution is consistent with the theoretical expectation within a factor 2. This difference can be explained by the uncertainty of the atmospheric muon flux close to horizon at the detector depth <sup>?)</sup>, and by uncertainties in the dead-time of individual channels. The highest multiplicity of hit channels (one event) is $`N_{hit}=24`$. Since no events with $`N_{hit}>24`$ are found in our data we can derive upper limits on the flux of high energy neutrinos which produce events with multiplicity $$N_{hit}>25.$$ (5) The effective volume of NT-96 for neutrino induced events depends only slightly on the value of the threshold multiplicity in condition (5). For the stronger conditions $`N_{hit}>27`$ and $`N_{hit}>29`$, the effective volume decreases by only 11% and 27%, respectively. The effective volume for neutrino produced events which fulfill conditions (4)-(5) was calculated as a function of neutrino energy and zenith angle $`\theta `$. The energy dependences of the effective volumes for isotropic electron and muon neutrinos are shown in Fig.7 (solid lines). Also shown are the effective volumes folded with the neutrino absorption probability in the Earth (dashed lines). The neutrino absorption in the Earth has been taken into account with suppression factor $`\mathrm{exp}(l(\mathrm{\Omega })/l_{tot})`$, where $`l(\mathrm{\Omega })`$ is the neutrino path length through the Earth in direction $`\mathrm{\Omega }`$ and $`l_{tot}^1=N_A\rho _{Earth}(\sigma _{CC}+\sigma _{NC})`$ according to <sup>?,?)</sup>. ### 3.4 The limits to the diffuse neutrino flux The number of events due to neutrino flux $`\mathrm{\Phi }_\nu `$ and processes (1) and (2) is given by $$N_\nu =Tϵ𝑑\mathrm{\Omega }\underset{\nu }{}𝑑EV_{eff}(\mathrm{\Omega },E)\underset{k}{}𝑑E_\nu \mathrm{\Phi }_\nu (E_\nu )N_A\rho _{H_2O}\frac{d\sigma _{\nu k}}{dE}$$ (6) where $`E_\nu `$ is the neutrino energy and $`E`$ \- the energy transferred to a shower or high energy muon. The index $`\nu `$ indicates the summation over neutrino types ($`\nu =\nu _\mu ,\overline{\nu _\mu },\nu _e`$) and $`k`$ over CC and NC interactions. $`N_A`$ is the Avogadro number, $`ϵ=0.9`$ \- the detector efficiency. Cross sections from R.Gandhi et al.(1996) have been used. The shape of the neutrino spectrum was assumed to behave like $`E^2`$ as typically expected for Fermi acceleration. In this case, 90% of expected events would be produced by neutrinos from the energy range $`10^4÷10^7`$GeV with the center of gravity around $`2\times 10^5`$GeV. Comparing the calculated rates with the upper limit to the actual number of events, 2.3 for 90% CL, and assuming the flavor ratios $`\mathrm{\Phi }_{\nu _\mu }=\mathrm{\Phi }_{\overline{\nu _\mu }}=\mathrm{\Phi }_{\nu _e}`$ due to photo-meson production of $`\pi ^+`$ followed by the decay $`\pi ^+\mu ^++\nu _\mu e^++\nu _e+\overline{\nu _\mu }+\nu _\mu `$ for extraterrestrial sources <sup>?,?)</sup>, we obtain the following upper limit to the diffuse neutrino flux: $$\frac{d\mathrm{\Phi }_\nu }{dE}E^2<1.43\times 10^5\text{cm}^2\text{s}^1\text{sr}^1\text{GeV}.$$ (7) New theoretical upper bounds to the intensity of high-energy neutrinos from extraterrestrial sources have been presented recently <sup>?,?)</sup>. These upper bounds as well as our limit and limits obtained by DUMAND <sup>?)</sup>, AMANDA <sup>?)</sup>, EAS-TOP <sup>?)</sup> and FREJUS <sup>?)</sup> experiments are shown in Fig.8. Also, the atmospheric neutrino fluxes <sup>?)</sup> from horizontal and vertical directions (upper and lower curves, respectively) are presented. For resonant process (3) the event number is given by: $$N_{\overline{\nu _e}}=Tϵ𝑑\mathrm{\Omega }𝑑EV_{eff}(\mathrm{\Omega },E)\underset{(M_w2\mathrm{\Gamma }_w)^2/2m_e}{\overset{(M_w+2\mathrm{\Gamma }_w)^2/2m_e}{}}𝑑E_\nu \mathrm{\Phi }_{\overline{\nu _e}}(E_\nu )\frac{10}{18}N_A\rho _{H_2O}\frac{d\sigma _{\overline{\nu _e},e}}{dE}$$ (8) $$M_w=80.22\text{GeV},\mathrm{\Gamma }_w=2.08\text{GeV}.$$ Our 90% CL limit at the W resonance energy is: $$\frac{d\mathrm{\Phi }_{\overline{\nu }}}{dE_{\overline{\nu }}}3.6\times 10^{18}\text{cm}^2\text{s}^1\text{sr}^1\text{GeV}^1.$$ (9) This limit lies between limits obtained by DUMAND ($`1.1\times 10^{18}`$ cm<sup>-2</sup>s<sup>-1</sup>sr<sup>-1</sup>GeV<sup>-1</sup>) and EAS-TOP ($`7.6\times 10^{18}`$cm<sup>-2</sup>s<sup>-1</sup>sr<sup>-1</sup>GeV<sup>-1</sup>). The new limits (10) and (12) have been obtained with the underwater detector NT-96. We hope that the analysis of 3 years data taking with NT-200 <sup>?,?)</sup> will allow us to lower this limit substantially. ### 3.5 Search for Fast Monopoles ($`\beta >0.75`$) Fast bare monopoles with unit magnetic Dirac charge and velocities greater than the Cherenkov threshold in water ($`\beta =v/c>0.75`$) are promising survey objects for underwater neutrino telescopes. For a given velocity $`\beta `$ the monopole Cherenkov radiation exceeds that of a relativistic muon by a factor $`(gn/e)^2=8.3\times 10^3`$ ($`n=1.33`$ \- index of refraction for water) <sup>?,?)</sup>. Therefore fast monopoles with $`\beta 0.8`$ can be detected up to distances $`55`$ m $`÷`$ $`85`$ m which corresponds to effective areas of (1–3)$`\times 10^4`$ m<sup>2</sup>. The natural way for fast monopole detection is based on the selection of events with high multiplicity of hits. In order to reduce the background from downward atmospheric muons we restrict ourself to monopoles coming from the lower hemisphere. Two independent approaches have been used for selection of upward monopole candidates from the 70 days of NT-96 data. The first one is similar to the method which was applied to upward moving muons (see subsection 3.2), with an additional cut $`N_{hit}>25`$ on the hit multiplicity. The second one cuts on the value of space-time correlation, followed by a cut $`N_{hit}>35`$ on the hit multiplicity. The upper limits on the monopole flux obtained with the two different methods coincide within errors. The same type of analysis was applied to the data taken during $`0.42`$ years lifetime with the neutrino telescope NT-36 <sup>?)</sup>. The combined $`90\%`$ C.L. upper limit obtained by the Baikal experiment for an isotropic flux of bare fast magnetic monopoles is shown in Fig.9, together with the best limits from underground experiments Soudan2, KGF, MACRO and Ohya <sup>?,?,?,?)</sup>. ## 4 Limnology Besides physics goals NT-200 can be used as a powerful tool to monitor water parameters. The array permanently records phototubes counting rates, and periodically parameters like optical transmission at various wavelengths, temperature, conductivity, pressure, and sound velocity. All these data complement traditional limnological studies are of importance to get a comprehensive understanding of processes occurring in the lake. Just for illustration we show OMs counting rate variations at various time scales recorded with NT-36 in 1993/94. Counting rates of individual OMs as well as individual channels (coincidence rate of a pair OMs) are dominated by water luminescence. Fig.10 presents the counting rate over 2 years and compares it to the bacteria concentration measured at distance of 50 km to the NT-200 site, at 10 m depth below surface. In August/September we observe an increase of the luminosity to extremely high levels. The changes of the counting rate of channels are not reflected in the muon trigger rate, since the muon trigger is dominated by atmospheric muons, with negligible contribution by random hits (water luminescence or dark current pulses). This is demonstrated in fig.11 on a shorter time scale, for a time interval of marked changes of individual channel counting rates following a severe storm at August 3rd, 1993, which had washed a lot of water from a nearby rivers and crooks to the lake. Fig.12 shows a short period of about 8 hours when the counting rates sequentially increased, starting with highest OMs and ending with the lowest ones along the string. From the time shift of the 3 curves a vertical current of $`2.3`$ cm$``$s<sup>-1</sup> is deduced. This is remarkable since the vertical velocity of water renewal is considered to be most intensively, is only $`0.2÷0.3`$ cm$``$s<sup>-1</sup>. ## 5 Conclusions and Outlook The results obtained with intermediate detector stages show the capability of Baikal Neutrino Telescope to search for the wide variety of phenomena in neutrino astrophysics, cosmic ray physics and particle physics. The first atmospheric neutrinos have been identified. Also limits on the fluxes of magnetic monopoles, diffuse flux of very high energy neutrinos as well as of neutrinos from WIMP annihilation in the center of the Earth have been derived. In the following years, NT-200 will be operated as a neutrino telescope with an effective area between 1000 and 5000 m<sup>2</sup>, depending on the energy, and will investigate atmospheric neutrino spectra above 10 GeV. NT-200 can be used to search for neutrinos from WIMP annihilation and for magnetic monopoles and high energy extraterrestrial neutrinos. It will also be a unique environmental laboratory to study water processes in Lake Baikal. Apart from its own goals, NT-200 is regarded to be a prototype for the development a large scale telescope of next generation with an effective area of several $`10^4`$ m<sup>2</sup>. The basic design of such a detector is under discussion at present. This work was supported by the Russian Ministry of Research,the German Ministry of Education and Research and the Russian Fund of Fundamental Research ( grants 99-02-18373a, 97-02-17935, 99-02-31006k, 97-02-96589 and 97-05-96466). REFERENCES
warning/0001/hep-ph0001106.html
ar5iv
text
# Separating Different Models from Measuring 𝛼,𝛽,𝛾 ## I Introduction In the standard model(SM) with $`SU(2)_L\times U(1)_Y`$ gauge symmetry and three generation of fermions. The only source of CP asymmetry is a non-zero complex phase in the Cabibbo-Kobayash-Maskawa(CKM) matrix. Although SM has been proved to be very successful in phenomenology, its accommodation of CP violation through complex CKM matrix elements has not been seriously tested experimentally. At present, CP violation is one of the least understood issues in particle physics, and is very promising in the search of indications of new physics. Assuming unitarity of CKM matrix, the phase information of the matrix can be displayed elegantly by a set of triangles, called Unitarity triangles(UTs) . The central topic is therefore the determination of the angles of those triangles. In the past years, much efforts have been made in the neutral-Kaon system as well as in $`B_d^0`$ and $`B_s^0`$ mixings. However, due to large theoretical uncertainties, our current knowledge of those angles is still very poor. It is expected that in the up coming B-factories, the measurements of time dependent CP asymmetries in $`B^0`$ decays into CP eigenstates will greatly reduce the hadronic uncertainties and obtain the precise value of the angles of UT with $`b`$ and $`d`$ quarks. If those angles are precisely determined, any deviation from the SM predictions will clearly signal new physics beyond the SM. Supposing all the angles ($`\alpha ,\beta ,\gamma `$) are completely determined through independent measurements, following the analysis in reference , there are three distinct ways in which new physics can show up in the measurements of CP asymmetry, they are: 1)$`\alpha +\beta +\gamma \pi `$, 2)$`\alpha +\beta +\gamma =\pi `$, but the value of $`\alpha ,\beta `$ and $`\gamma `$ do not agree with the SM predictions. 3) $`\alpha +\beta +\gamma =\pi `$, $`\alpha ,\beta `$ and $`\gamma `$ are consistent with the SM, but measurement of the angles are inconsistent with the measurements of the side of the UT. If any one of these three cases really happens in the future B- factory, the new physics will be established. However, this only tells us that new physics exactly exists. We still don’t know what kind of new physics is responsible, since there are variety of models of new physics which can affect the value of angles in the same way. It is of great importance to distinguish different models of new physics from the experiment. The problem of distinguishing various models of new physics has been discussed in . It may involve many models of new physics such as suppersymmetric models,muti-higgs without flavor changing neutral scalars interactions or general two-higgs-doublet model with flavor changing neutral scalars interactions (S2HDM) , left-right symmetric model, Z-mediated flavor changing neutral currents(FCNC) and fourth generation. The conclusion is that by comparing their contributions to $`B^0\overline{B}^0`$ mixings and rare leptonic $`B`$ decays, these models can be partially distinguished. If new physics is founded to be the case 1) or 2) as mentioned above. This would indicate that new physics is probably to be the S2HDM, fourth-generation or Z-mediated FCNC. If new physics is founded through the case 3), the new physics is likely to be two-Higgs doublet model without flavor changing neutral scalar interactions (such as Model 1 or Model 2) or minimum suppersymmetry model. This method is very useful, but still not sufficient to distinguish each of the models especially when several models have similar effect on the extraction of those unitarity angles. In this paper, We show that the combination of analysis on $`B^0\overline{B}^0`$ mixing and $`B`$-decay amplitude via time dependent measurement of CP asymmetry is also an efficient way in separating different models. As an example , the top quark two-Higgs doublet model(T2HDM) recently discussed in can be distinguished not only from a large number of model without CP asymmetry in $`B`$-decay amplitude, but also from the S2HDM by their different behavior in $`B^0\overline{B}^0`$ mixing. The paper is organized as follows: In section II , we present the basic formulas on distinguishing different models of new physics by considering their difference in $`B^0\overline{B}^0`$ mixing and $`B`$-decay amplitude, some features of the S2HDM and T2HDM as well as their influences on the determination of angles $`\alpha ,\beta `$ and $`\gamma `$ are discussed in section III. The conclusions are presented in section IV. ## II basic formulas From the theoretical point of view, there are two basic ways in which new physics can enter the extraction of angles $`\alpha ,\beta `$ and $`\gamma `$. One way is via $`B^0\overline{B}^0`$ mixings, the other is via $`B`$ decay amplitudes which is mainly through hadronic penguin diagrams, but in some models it may be also through tree diagrams, such as the S2HDM and T2HDM. That will be discussed in detail below. If $`B^0\overline{B}^0`$ mixing is affected by new physics, for example, from additional heavy particles in the loop instead of $`W`$-boson, the angle $`\beta `$ measured in the process $`B_d^0J/\psi K_S`$ can be largely modified. In the SM the time dependent asymmetry is given as follows: $$Im\lambda =\left(\frac{q}{p}\right)_{B_d}\left(\frac{\overline{A}}{A}\right)\left(\frac{p}{q}\right)_K=\mathrm{sin}2\beta ,$$ (1) where the term in the first bracket is from $`B^0\overline{B}^0`$ mixing which has the value of $`V_{td}V_{tb}^{}/V_{td}^{}V_{tb}`$ in the SM, $`A(\overline{A})`$ is the amplitude of $`bc(\overline{c}s)(\overline{b}\overline{c}(c\overline{s}))`$ subprocess, the term in the last bracket is from $`K^0\overline{K}^0`$ mixing since $`K_S`$ is involved in the final state. Without losing generality, the new physics can affect all these three quantities. Let us denote $`\varphi _{mix}^{B_d},\varphi _A`$ and $`\varphi _{mix}^K`$ the new phase from new physics in $`B_d^0`$ mixing, $`B`$-decay amplitude, and $`K^0`$ mixing respectively. Then the experiment will measure $`\beta _{exp}`$ instead of $`\beta _{SM}`$ as: $$\beta _{exp}=\beta _{SM}+\varphi _{mix}^{B_d}+\varphi _A(bc)+\varphi _{mix}^K$$ (2) where the process $`bc`$ indicated in the bracket is a tree level transition. The angle $`\alpha `$ can be measured from $`B_d^0`$ decay to $`\pi ^+\pi ^{}`$. In this channel there are also contributions from penguin diagrams with different strong phase, this can be eliminated by using the isospin analysis. As it was pointed out in and also emphasized by many authors, if the angle $`\alpha `$ is measured through the decay channel $`B_d^0\pi ^+\pi ^{}`$, then the new physics effect will give contributions with an opposite sign as follows: $$\alpha _{exp}=\pi \beta _{SM}\gamma _{SM}\varphi _{mix}^{B_d}\varphi _A(bu)$$ (3) Here the process is ($`bu`$) as $`B_d^0\pi ^+\pi ^{}`$ is dominated by $`bu`$ transition. Consequently, the new phase $`\varphi _{mix}^{B_d}`$ in $`B^0\overline{B}^0`$ mixing $`cancels`$ each other in the sum $`\alpha _{exp}+\beta _{exp}`$. When the new phase from amplitudes and $`K`$-meson mixing are negligible small( these happens in many models ), the sum $`\alpha +\beta `$ will remain unchanged. If the angle $`\gamma `$ is determined through charged $`B`$ decay $`B^\pm DK^\pm `$, since $`B^0`$-mixing is absent and the FCNC will not be involved in this channel, its value can hardly be modified by new physics, so $`\gamma _{exp}`$ is likely to be unchanged and equals $`\gamma _{SM}`$. Thus the sum $$\alpha _{exp}+\beta _{exp}+\gamma _{exp}=\pi $$ (4) still holds as in the case of SM. The new phase $`\varphi _{mix}^K`$ in $`K`$-mixing is often thought to be small, this is because the extremely small values of $`\mathrm{\Delta }m_K`$ and $`ϵ_K`$ impose a very strong constraint on the contributions to $`K^0\overline{K}^0`$ mixing from new physics. Thus as a consequence, the new physics can not produce a relative large value of $`\varphi _{mix}^K`$. It is a possibility that by observing the violation of equation: $$Im\lambda (B_dD^+D^{})=Im\lambda (B_dJ/\psi K_S)$$ (5) or $$Im\lambda (B_dD^+D^{})=Im\lambda (B_d\varphi K_S)$$ (6) One is able to probe the new physics in $`K^0\overline{K}^0`$ mixing . In the following discussion we always make the assumption that the new phase in $`K^0`$ mixing is negligible. Although the effect of decay amplitude $`\varphi _A`$ is always thought to be small, its importance on signaling new physics should not be neglected. As being stressed in reference , the effects of new physics in decay amplitudes are manifestly non-universal, because they strongly depend on the specific process and decay channel under consideration. On the other hand, the effects on $`B^0\overline{B}^0`$ mixing are almost insensitive to the decay modes. Since in general $`\varphi _A(bc)\varphi _A(bu)`$ , in the condition that $`\varphi _{mix}^K`$ is zero, the equation(4) becomes $$\alpha _{exp}+\beta _{exp}+\gamma _{exp}=\pi +\varphi _A(bc)\varphi _A(bu).$$ (7) this will be a clear signal of new physics from decay amplitude. By considering whether the equation(4) holds, the models of new physics can be cataloged into two classes, i.e. models with or without new phase in decay amplitude. A large number of models such as the 2HDM of types I and II, left-right symmetric, and the minimum suppersymmetric model, fourth generation and Z-mediated flavor changing neutral currents(FCNC) fall into the first class, whereas the S2HDM and T2HDM as well as some other models fall into the second class. Therefore the models can be partially distinguished in this way. Another useful information is the new phase in $`B^0\overline{B}^0`$ mixing. However, due to their cancellation in the sum $`\alpha +\beta `$ it can not be extracted directly. If one looks at the $`B_d^0`$ decay to CP eigenstates such as $`J/\psi K_S,D^+D^{}`$ or $`\varphi K_S`$ the new phases from $`B^0`$-mixing and from $`B`$ decay amplitude may mix with each other, and the final result is the sum of these two kind of contributions. Thus the extraction of pure new phase from $`B^0\overline{B}^0`$ mixing in $`B_d^0J/\psi K_S`$ becomes difficult. This situation can be simplified if one of the decay amplitudes $`\varphi _A(bc)`$ and $`\varphi _A(bu)`$ is negligible small. For example, $`\varphi _A(bu)0`$ while $`\varphi _A(bc)`$ is obviously non-zero ( this happens in the case of the S2HDM and T2HDM which is under consideration of this paper ), thus equation(7) becomes $$\alpha _{exp}+\beta _{exp}+\gamma _{exp}=\pi +\varphi _A(bc).$$ (8) Thus $`\varphi _A(bc)`$ can be easily obtained by measuring the sum of $`\alpha ,\beta ,\text{and }\gamma `$. Substituting the value of $`\varphi _A(bc)`$ into equation(2), the phase $`\varphi _{mix}^{B_d}`$ can be fixed. In doing this, we need to know the SM prediction of $`\beta _{SM}`$ since many decay channels can be seriously polluted by new physics, one should carefully choose some processes which are not likely to be modified. One way is to use the value of $`|V_{cb}|,|V_{ub}|/|V_{cb}|`$ from semileptonic $`B`$ decays $`bc(u)l\overline{\nu _l}`$ and $`\gamma `$ from $`B^\pm DK^\pm `$. These three quantities correspond to two sides and one angle between the two sides in UT. Thus the whole triangle including angle $`\beta `$ can be completely determined. The shortcoming here is that the prediction of $`|V_{ub}|/|V_{cb}|`$ is model dependent and suffer a large theoretical uncertainties. An alternative way to extract $`\beta `$ is via rare $`K`$ decay $`K\pi \nu \overline{\nu }`$. The branching ratio of decay $`K^+\pi ^+\nu \overline{\nu }`$ is given as follows: $$B(K^+\pi ^+\nu \overline{\nu })=\kappa \left[\left(\frac{Im\lambda _t}{\lambda ^5}X(x_t)\right)^2+\left(\frac{Re\lambda _c}{\lambda }P_0(K^+)+\frac{Re\lambda _t}{\lambda ^5}X(x_t)\right)^2\right]$$ (9) with $$\kappa =\frac{3\alpha ^2B(k^+\pi ^0e^+\nu )}{2\pi ^2\mathrm{sin}^4\theta _W}\lambda ^8=4.64\times 10^{11}$$ (10) where $`X(x_i)`$ is an integral function given in , $`x_t=m_t^2/m_W^2,\lambda _i=V_{is}^{}V_{id}`$ and $`\lambda =|V_{us}|0.22`$. The function $`P_0(K^+)`$ has the form $`P_0(K^+)=(2X_{NL}^e/3+X_{NL}^\tau /3)/\lambda ^4`$. By combining the branching ratio of $`K_L\pi ^0\nu \overline{\nu }`$: $`B(K_L\pi ^0\nu \overline{\nu })`$ $`=`$ $`\kappa _L\left({\displaystyle \frac{Im\lambda _t}{\lambda ^5}}X(x_t)\right)^2`$ (11) $`\kappa _L=\kappa {\displaystyle \frac{\tau (K_L)}{\tau (K^+)}}`$ $`=`$ $`1.94\times 10^{10}`$ (12) one can find: $$Im\lambda _t=\lambda ^5\frac{\sqrt{B_2}}{X(x_t)}Re\lambda _t=\lambda ^5\frac{\frac{Re\lambda _c}{\lambda }P_0(K^+)+\sqrt{B_1B_2}}{X(x_t)}$$ (13) where $`B_1`$ and $`B_2`$ are the reduced branching ratios with $`B_1=B(K^+\pi ^+\nu \overline{\nu })/4.64\times 10^{11}`$ and $`B_2=B(K^+\pi ^0\nu \overline{\nu })/1.94\times 10^{10}`$. Using the standard parameterization of the CKM matrix, the angle $`\beta `$ can be determined by $$\mathrm{sin}2\beta =\frac{2r_s}{1+r_s^2}$$ (14) with $`r_s=(1\overline{\rho })/\overline{\eta }`$. The parameter $`\overline{\rho }`$ and $`\overline{\eta }`$ is given as follows: $`\overline{\rho }`$ $`=`$ $`{\displaystyle \frac{\sqrt{1+4s_{12}c_{12}Re\lambda _t/s_{23}^2(2s_{12}c_{12}Im\lambda _t/s_{23}^2)^2}1+2s_{12}^2}{2c_{23}^2s_{12}^2}}`$ (15) $`\overline{\eta }`$ $`=`$ $`{\displaystyle \frac{c_{12}Im\lambda _t}{s_{12}c_{23}^2s_{23}^2}}`$ (16) It is well known that the rare decays $`K^+\pi ^+\nu \overline{\nu }`$ and $`K_L\pi ^0\nu \overline{\nu }`$ can be calculated with smaller theoretical uncertainties. These uncertainties can be further reduced in the next-to-leading order QCD corrections . As a result, the measurement of both two decays with an error of $`\pm 10\%`$ will yield $`\mathrm{sin}2\beta `$ with an accuracy comparable to the determination from CP asymmetry in $`B`$-decays prior to LHC. Here we emphasize that these channels are not likely to be affected by new physics models, especially, the models with FCNC. This is because the couplings between fermions and additional scalars which often present in the models of new physics are proportional to the fermion mass, the decay involving leptons in the final states will greatly suppress the tree level contributions from those scalars. Although there may be significant new physics contributions to $`Z`$-penguin diagrams, the angle $`\beta `$ will remain unchanged if the new physics do not carry additional new phase. This happens in many models with additional Higgs bosons, such as 2HDM of type I, and type II, minimum supersymmetric models, et.al. If $`\beta _{SM}`$ is extracted in this way, it is then possible to study the behavior of different models in $`B^0\overline{B^0}`$ mixing independent of $`B`$-decay amplitudes. In general, different models have different behavior in $`B^0`$ mixings and decay amplitudes. It is therefore possible to identify the models by combining the analysis of these two aspects. Following this strategy, the T2HDM can be distinguished not only from those models without new phase in $`B`$ decay amplitudes, but also from the S2HDM by its absence of contribution in $`B^0`$ mixing. This will be further discussed in the next section. ## III On S2HDM and T2HDM Let us briefly present some important prospects of T2HDM. The Lagrangian of T2HDM is as follows: $$_Y=\overline{L}_L\varphi _1El_R\overline{Q}_L\varphi _1Fd_R\overline{Q}_L\stackrel{~}{\varphi }_1GI^{(1)}u_R\overline{Q}_L\stackrel{~}{\varphi }_2GI^{(2)}u_R+h.c,$$ (17) Where $`\overline{L}_L`$ and $`\overline{Q}_L`$ are the ordinary left-handed lepton and quark doublets, $`u_R`$ and $`d_R`$ are right-handed singlet quarks, $`\varphi _1`$ and $`\varphi _2`$ are two Higgs doublets with $`\stackrel{~}{\varphi }_i=i\sigma ^2\varphi _i^{}`$ and $`E,F,G`$ are Yukawa coupling matrix. $`I^{(1)}`$ and $`I^{(2)}`$ are two diagonal matrix with $`I_{ij}^{(1)}=\delta _{ij}(i,j=1,2)`$ and $`I_{ij}^{(2)}=\delta _{ij}(i=j=3)`$. Comparing with other quarks, the top quark is in a special status in this model, i.e. Only $`\varphi _2`$ couples to $`t_R`$. Let the vacuum expectation value(VEV) of two Higgs fields to be $`v_1/\sqrt{2}`$ and $`v_2e^{i\delta }/\sqrt{2}`$ respectively. If we choose the ratio between two VEVs $`\mathrm{tan}\beta =|v_2|/|v_1|`$ to be large ($`\mathrm{tan}\beta `$ is close to $`m_t/m_b`$), the large mass of top quark can be naturally explained. That is the motivation of proposing this model. The charged quark-Higgs Yukawa interaction in this model reads $`^C=`$ $``$ $`2\sqrt{2}G_F[\overline{u}_L^iV_{ij}m_{d_j}d_R^j\mathrm{tan}\beta `$ (18) $`+`$ $`\overline{u}_R^imu_iV_{ij}d_R^j\mathrm{tan}\beta +\overline{u}_R^i\mathrm{\Sigma }_{ij^{}}^{}V_{j^{}j}d_L^j(\mathrm{tan}\beta +\mathrm{cot}\beta )]H^++H.c,`$ (19) where $`m_i`$ are the quark mass eigenstates, V is the usual CKM matrix. The matrix $`\mathrm{\Sigma }`$ can be parameterized as follows: $$\mathrm{\Sigma }=\left(\begin{array}{ccc}0& 0& 0\\ 0& m_cϵ_{ct}^2|\xi |^2& m_cϵ_{ct}\xi ^{}\sqrt{1|ϵ_{ct}|\xi |^2}\\ 0& m_c\xi \sqrt{1|ϵ_{ct}|\xi |^2}& m_t(1|ϵ_{ct}\xi |^2)\end{array}\right)$$ (20) Since there exist none zero off-diagonal elements in $`\mathrm{\Sigma }`$ matrix, this model may lead to flavor changing neutral currents, but only in up-type quarks. Therefore it will enhance the $`D^0\overline{D}^0`$ mixing provided that $`\mathrm{tan}\beta `$ is large. On the other hand, it has little effect on $`K^0\overline{K}^0`$ mixing and $`B^0\overline{B}^0`$ mixing since these mesons contain down-type quarks. Another distinct feature is that the off-diagonal element $`\mathrm{\Sigma }_{32}`$ can be quit large, thus the model generally has a very large couplings for the vertex $`\overline{b}cH^{}`$ or $`\overline{t}cH^0`$ . Let us turn to a brief discussion on the S2HDM. This model can be obtained if we abandon the discrete symmetry which is often imposed on the Lagrangian of two Higgs doublet model and replac it with an approximate global U(1) family symmetry. The point is that the smallness of the off-diagonal terms in the CKM matrix suggests that violation of flavor symmetry are specified by small parameters. It then turns out that reasonable choices for these small parameters combined with the natural smallness of Higgs boson couplings allows one to meet the constraint on flavor changing neutral scalar exchange. Since there are no discrete symmetries, many new sources of CP asymmetry can arise from its lagrangian. Therefore, this model can affect the measurement of the angles of UT in many different ways. For example the possible large effects on $`B^0\overline{B}^0`$ mixing, weak transition $`tc`$ and large CP asymmetry in $`bs\gamma `$ have been investigated. The Lagrangian of S2HDM has the form: $$_Y=\overline{Q}^i\mathrm{\Gamma }_{1,ij}^UU_{R_j}\varphi _1+\overline{Q}^i\mathrm{\Gamma }_{1,ij}^DD_{R_j}\stackrel{~}{\varphi _1}+\overline{Q}^i\mathrm{\Gamma }_{2,ij}^UU_{R_j}\varphi _2+\overline{Q}^i\mathrm{\Gamma }_{2,ij}^DD_{R_j}\stackrel{~}{\varphi _2}+h.c.$$ (21) After a rotation into quark mass eigenstates, it can be rewritten as: $$_Y=(L_1+L_2)(\sqrt{2}G_F)^{1/2}$$ (22) with $`_1`$ $`=`$ $`\sqrt{2}(H^+{\displaystyle \underset{i,j}{\overset{3}{}}}\xi _{d_j}m_{d_j}V_{ij}\overline{u}_L^id_R^jH^{}{\displaystyle \underset{i,j}{\overset{3}{}}}\xi _{u_j}m_{u_j}V_{ij}^{}\overline{d}_L^iu_R^j)`$ (25) $`+H^0{\displaystyle \underset{i}{\overset{3}{}}}(m_{u_i}\overline{u}_L^iu_R^i+m_{d_i}\overline{d}_L^id_R^i)`$ $`+(R+iI){\displaystyle \underset{i}{\overset{3}{}}}\xi _{d_i}m_{d_i}\overline{d}_L^id_R^i+(RiI){\displaystyle \underset{i}{\overset{3}{}}}\xi _{u_i}m_{u_i}\overline{u}_L^iu_R^i+H.c.`$ $`_2`$ $`=`$ $`\sqrt{2}(H^+{\displaystyle \underset{i,j^{}j}{\overset{3}{}}}V_{ij^{}}\mu _{j^{}j}^d\overline{u}_L^id_R^jH^{}{\displaystyle \underset{i,j^{}j}{\overset{3}{}}}V_{ij^{}}^{}\mu _{j^{}j}^u\overline{d}_L^iu_R^j)`$ (27) $`+(R+iI){\displaystyle \underset{ij}{\overset{3}{}}}\mu _{ij}^d\overline{d}_L^id_R^j+(RiI){\displaystyle \underset{ij}{\overset{3}{}}}\mu _{ij}^u\overline{u}_L^iu_R^j+H.c.`$ Where the factors $`\xi _{d_j}m_{d_j}`$ arise primarily from diagonal elements of $`\mathrm{\Gamma }_1`$ and $`\mathrm{\Gamma }_2`$ whereas the factors $`\mu _{jj^{}}^d`$ arise from the small off-diagonal elements. By abandoning the discrete symmetry, this model obtains rich sources of CP violation. They can be classified into four major types: (1) The induced CKM matrix. (2) The phases in the factors $`\xi _{f_i}`$ provide CP violation in the charged-Higgs exchange processes, which are independent of the CKM phase. (3) The phases in the factors $`\mu _{ij}^f`$. These yield CP violation in flavor changing neutral scalar interaction. (4) the phase from the mixing matrix of the three neutral Higgs scalars Although these two models have some similar behavior in CP asymmetry, there still exist several subtle differences between them. First, in the T2HDM the coupling between quarks and Higgs boson is determined by only parameters $`\mathrm{tan}\beta `$ and $`\xi `$. On the other hand such couplings in the S2HDM are flavor dependent. So that the latter has more freedom in fitting the experimental data. Second, there is no complex phase in the diagonal Yukawa couplings in T2HDM. This means that there is no CP asymmetry from charged Higgs exchange in $`tb`$ transition. So it will result in a small CP asymmetry in the decay $`bs\gamma `$ , which is of the order less than $`10^2`$. On the contrary in the case of S2HDM, this effect could be larger. Third, although in T2HDM the non-zero complex elements in $`\mathrm{\Sigma }`$ matrix can lead to FCNC, its effect is constrained only in up-type quarks, there is no FCNC between down-type quarks. As a result, $`K^0\overline{K}^0`$ and $`B^0\overline{B}^0`$ mixings can not be seriously modified by this model. But they could receive contributions in the S2HDM. Let us investigate their new physics effects on the determination of angles $`\alpha ,\beta ,\gamma `$, respectively . The ’gold-plated’ channel for determining angle $`\beta `$ is decay $`B_d^0J/\psi K_S`$, which is dominated by tree level $`bc`$ process. In both S2HDM and T2HDM, a considerable contribution from decay amplitude can arise from $`bc`$ transition . The reason is that the off-diagonal element $`\mathrm{\Sigma }_{32}`$ can relatively large in T2HDM. Moreover one can see from equation(3.3) that the CKM matrix elements associated with $`\mathrm{\Sigma }_{32}`$ is $`V_{tb}`$ rather than $`V_{cb}`$. This can contribute to an additional enhancement factor of $`|V_{tb}/V_{cb}|25`$. The effective Lagrangian at tree level $`bc\overline{c}s`$ has the form: $$L_{eff}=2\sqrt{2}G_FV_{cb}V_{cs}^{}\left[\overline{c}_L\gamma _\mu b_L\overline{s}_L\gamma ^\mu c_L+2\zeta e^{i\delta }\overline{c}_Rb_L\overline{s}_Lc_R\right],$$ (28) where $$\zeta e^{i\delta }=\{\begin{array}{cc}(1/2)(V_{tb}/V_{cb})(m_c\mathrm{tan}\beta /m_H)^2\xi ^{}\hfill & \text{for T2HDM}\\ & \\ (1/2)(V_{tb}/V_{cb})(\mu _{32}^u/m_H)^2\hfill & \text{for S2HDM}\end{array}$$ (29) Using the formalism in references. The decay amplitude of $`BJ/\psi K_S`$ can be written as $`A=A_{SM}[1\zeta e^{i\delta }]`$, where $`A_{SM}`$ denotes the amplitude in SM. If factorization holds there is no relative strong phase between $`W`$-and Charged-Higgs exchange process. So $`B`$ and $`\overline{B}`$ decay only differ by a CP-violating phase. Their ratio is given by: $`\overline{A}/A=(\overline{A}_{SM})/A_{SM}e^{2i\varphi _A}`$, where $`\varphi _A=\mathrm{tan}^1(\zeta \mathrm{sin}\delta /(1\zeta \mathrm{cos}\delta ))`$ is the correction to SM from charged Higgs exchange. Thus the time dependent asymmetry will measure $`\beta _{SM}+\varphi _A`$ rather than $`\beta _{SM}`$ In the case of S2HDM, the new phase from amplitude can be obtained by simply replace the expression of $`\zeta e^{i\delta }`$ in T2HDM in equation(29). However, the situation here is more complicate since there are additional contributions to $`B^0\overline{B}^0`$ mixing, which will largely change the value of angle $`\beta `$. The new contributions come from the couplings $`\xi _i`$ and $`\mu _{ij}`$. They are in general complex. Although the value of $`\mu _{ij}`$ can be constrained from the measurement in $`x_d\mathrm{\Delta }m_B/\mathrm{\Gamma }_B`$, due to the large uncertainties of $`|V_{td}|`$ , $`\varphi _{mix}^B`$ can still be rather large. The decay $`B_d^0\pi ^+\pi ^{}`$ is thought to be a good channel for the extraction of angle $`\alpha `$. Since $`B_d^0\pi ^+\pi ^{}`$ decay is dominated by $`bu(\overline{u}d)`$ tree level process. The additional contribution from charged Higgs boson is proportional to the $`d`$-quark mass $`m_d`$ which is negligibly small. There are also contributions from charged Higgs loop in penguin diagrams. By using the isospin analysis, those effects can be eliminated. Thus there are no additional new phases in $`B_d^0\pi ^+\pi ^{}`$ decay amplitude. In T2HDM, due to the absence of new phase in $`B^0\overline{B}^0`$ mixing, this measurement of angle $`\alpha `$ will give $`\alpha _{exp}=\alpha _{SM}`$, however in the S2HDM, as has been discussed in the previous section, it will give $`\alpha _{exp}=\alpha _{SM}\varphi _{mix}^B`$. Finally let us consider the measurements of angle $`\gamma `$. As being proposed in , $`\gamma `$ can be extracted from charged $`B`$ decay $`B^\pm DK^\pm `$. By an independent measurement of six amplitude $`B^+D^0K^+,B^+\overline{D}^0K^+,B^+D_{CP}^0K^+,B^{}D^0K^{},B^{}\overline{D}^0K^{},B^{}D_{CP}^0K^{}`$ the angle $`\gamma `$ can be in principle determined. Where $`D_{CP}^0=\frac{\sqrt{2}}{2}(D^0+\overline{D}^0)`$ is the CP even eigenstate. In the SM the relations of these six amplitudes are : $`\sqrt{2}A(B^+D_1^0K^+)`$ $`=`$ $`A(B^+D^0K^+)+A(B^+\overline{D}^0K^+),`$ (30) $`\sqrt{2}A(B^{}D_1^0K^{})`$ $`=`$ $`A(B^{}D^0K^{})+A(B^{}\overline{D}^0K^{}),`$ (31) and $`A(B^+\overline{D}^0K^+)=A(B^{}D^0K^{}),`$ (32) $`A(B^+D^0K^+)=A(B^{}\overline{D}^0K^{}).`$ (33) with $`|A(B^+D_1^0K^+)||A(B^{}D_1^0K^{})|`$ these relations are illustrated in Fig. 1. If the magnitudes of the amplitudes can be measured experimentally, one can then extract the angle $`\gamma `$. In both S2HDM and T2HDM, only $`bc`$ process could be modified considerably by charged Higgs exchange. This implies that the two amplitudes $`A(B^+D^0K^+)`$ and $`A(B^+D^0K^+)`$ which are dominated by the $`bu`$ transitions will remain unchanged and the angle between them is still given by $`2\gamma `$. However, the new phase can contribute to $`A(B^+\overline{D}^0K^+)`$ and $`A(B^+\overline{D}^0K^+)`$ since they are dominated by the $`bc`$ at tree level subprocess. The relation of (32) will be modified to be $$A(B^+\overline{D}^0K^+)=e^{2i\varphi _A}A(B^{}D^0K^{})$$ (34) As it is shown in Fig.1, if $`\varphi _A`$ can be extracted from equation(8), then the angle $`\gamma `$ can be obtained by measuring those six amplitudes. The angle $`\gamma `$ determined in this way is equal to the one in the SM, i.e. $`\gamma _{exp}=\gamma _{SM}`$. In summary, the $`B_d^0J/\psi K_S`$ decay can be affected in both models. In the T2HDM, it is affected through the $`B`$ decay amplitudes. In the S2HDM , the new phase could arise from both $`B^0`$ mixing and decay amplitudes. In the extraction of $`\alpha `$, the effect of T2HDM is negligible, but the one of S2HD can contribute a new phase from decay amplitude through $`B^0`$ mixing. In the extraction of $`\gamma `$ , if the new phase from decay amplitude can be determined from equation(8) , the method of measuring $`\gamma `$ by combining the six amplitudes of $`B^\pm DK^\pm `$ still works and $`\gamma _{exp}`$ will be equal to $`\gamma _{SM}`$. Since $`\beta _{SM}`$ can be extracted through $`K\pi \nu \overline{\nu }`$, it is then possible to extract $`\varphi _{mix}^B`$ from (2.2) or (2.3). If $`\varphi _{mix}^B0`$ is observed in the future experiment, it implies that the T2HDM is disfavorable. ## IV conclusions In conclusion, the distinguishment of different new physics models is discussed . It has been seen that by comparing their different behaviors in $`B^0\overline{B}^0`$ mixing and $`B`$ decay amplitudes, these models can be partially separated. The distinguishments between T2HDM and S2HDM have been discussed in detail. The new physics effects in measuring UT angles $`\alpha ,\beta `$ and $`\gamma `$ from these two models have been examined. It has been seen that by measuring $`\alpha ,\beta ,\gamma `$ from $`B_d^0\pi ^+\pi ^{},K\pi \nu \overline{\nu }`$ and $`B^\pm DK^\pm `$ respectively, the new phase $`\varphi _{mix}^{B_d}`$ can be extracted. Since there is no contribution from $`B^0`$ mixing in T2HDM, if $`\varphi _{mix}^{B_d}0`$ from the future experiment is founded, the T2HDM will be excluded. The situation will be different in the S2HDM where a non-zero value of $`\varphi _{mix}^{B_d}`$ is allowed. This is because the T2HDM can be regarded as one of the special cases of S2HDM. This work was supported in part by the NSF of China under grant No. 19625514.
warning/0001/gr-qc0001032.html
ar5iv
text
# Analytic treatment of black-hole gravitational waves at the algebraically special frequency ## I Introduction The evolution of small perturbations of the Schwarzschild metric of a spherically symmetric black hole, such as generated by infalling matter or in the aftermath of a stellar collapse, is well known to be described in each angular-momentum sector $`\mathrm{}2`$ by a Klein–Gordon wave equation $$[_x^2+\omega ^2V(x)]\varphi (x,\omega )=0.$$ (1) Here, $`\varphi `$ is a scalar field (a combination of the metric-function perturbations), while the ‘tortoise coordinate’ is $$x=r+\mathrm{ln}(r1),$$ (2) where $`r`$ is the circumferential radius (so that $`1<r<\mathrm{}`$ maps to $`\mathrm{}<x<\mathrm{}`$), and $`c=G=2M=1`$ ($`M`$ is the black-hole mass). However, the axial sector is described by the Regge–Wheeler equation (RWE), which has the potential $$V(x)=(r1)\frac{2(n+1)r3}{r^4},$$ (3) while in the polar sector, the Zerilli equation (ZE) has $$\stackrel{~}{V}(x)=(r1)\frac{8n^2(n+1)r^3+12n^2r^2+18nr+9}{r^4(2nr+3)^2},$$ (4) where we defined $`n=\frac{1}{2}(\mathrm{}1)(\mathrm{}+2)`$. Clearly, one is now interested in the modes of these equations, i.e., eigenfunctions under either outgoing- ($`\varphi (x\pm \mathrm{},\omega )e^{i\omega |x|}`$; however, see below for precise definitions) or incoming-wave ($`\varphi (x,\omega )e^{i\omega |x|}`$) boundary conditions. Foremost among these are the quasinormal modes (QNMs, with outgoing boundary conditions), which constitute the excitation spectrum of the black hole. Besides being a matter of principle in deciding the stability of the Schwarzschild solution, the issue takes on exciting significance with the prospect of their experimental detection in the coming decade using instruments such as LIGO and VIRGO. QNMs also play a role in fully nonlinear numerical simulations, where they are observed to dominate the radiated signal at intermediate times after a violent event such as a stellar collapse or a black-hole collision (cf. the beginning of this section), so that numerical and perturbative studies are complementary. Besides QNMs, we will also consider total-transmission modes, incoming from the left (right) but outgoing to the other side, and denoted as TTM<sub>L</sub>s (TTM<sub>R</sub>s). The fourth type of mode, a normal mode (NM) or bound state with $`\varphi (x,\omega )e^{|\omega x|}`$, is ruled out since both $`V`$ and $`\stackrel{~}{V}`$ are purely repulsive. It has been noticed that the RWE and ZE have the same QNM spectrum, which a priori is by no means obvious given the very different forms for the potentials. This can be understood because solutions of the two equations are related by ‘intertwining’ or supersymmetry (SUSY) as $`\stackrel{~}{\varphi }=[d_x+W(x)]\varphi `$, with the superpotential $$W(x)=\frac{N}{2}+\frac{3(r1)}{r^2(2nr+3)},$$ (5) where $$N\frac{4n(n+1)}{3}=8(\begin{array}{c}\mathrm{}+2\\ 4\end{array}).$$ (6) Of particular interest are now the SUSY generator $$\xi _1(x)=\mathrm{exp}\left\{^x𝑑yW(y)\right\}$$ (7) (cf. Section II B for the notation) and its counterpart $`\stackrel{~}{\xi }_1^{}\xi _1^1`$, which generates the inverse transform. By the general theory of SUSY, these are solutions of the RWE and the ZE respectively, at the same eigenvalue, which in this case is yielded by (1) as $`\omega ^2=\mathrm{\Omega }^2`$ for the ‘algebraically special frequency’ $`\mathrm{\Omega }iN/2`$. Since the early study of ‘algebraically special’ black-hole perturbations, these have been associated with TTMs: Ref., dealing with the Schwarzschild case, states (below its last, unnumbered display) that ‘The radiation field (incoming) … vanishes.’ The analysis of was generalized to Kerr holes in, still on the level of general relativity, and subsequently Ref. gave the corresponding exact solutions of the RWE, ZE, and Teukolsky equations. Given the fact that consistently calls these solutions TTMs, it is remarkable that their solving the appropriate wave equations is checked explicitly, but not their obeying the corresponding boundary conditions. The situation was confounded when Ref. reported a QNM $`\mathrm{\Omega }^{}`$ of the RWE numerically very close to, but not at, $`\mathrm{\Omega }`$. For $`\mathrm{}=3`$, however, $`\mathrm{\Omega }^{}`$ is given with a nonzero real part, while the QNM $`\mathrm{\Omega }^{}`$ which would then be dictated by symmetry is not found; this seems to imply that the last 3–4 digits (the order of magnitude of the reported $`Re\mathrm{\Omega }^{}`$) given by are all insignificant, so that no conclusions can be drawn about possible coincidence of $`\mathrm{\Omega }`$ and $`\mathrm{\Omega }^{}`$ (if the latter does exist at all). In contrast, Ref. reports TTM<sub>R</sub>s of the RWE very close to $`\mathrm{\Omega }`$ for $`2\mathrm{}6`$, and suggests that these correspond to the special mode; again, if all the given decimals are significant, strictly speaking this suggestion is not warranted. Finally, Ref. finds that for $`\mathrm{}=2`$ and for each ‘magnetic quantum number’ $`2m2`$ the Kerr hole has a QNM (the ninth) which tends to $`\mathrm{\Omega }=4i`$ in the Schwarzschild limit $`a0`$ ($`a`$ is the black-hole rotation). However, Ref. cautions that these QNMs should disappear for $`a=0`$, since they cannot coexist at $`\mathrm{\Omega }`$ with the TTM furnished by the special mode. For each $`m`$, this disappearance is thought to occur by cancellation of the two modes (mirror images only for $`m=0`$) on each side of the negative imaginary axis (NIA) in the $`\omega `$-plane. Certainly this is conceivable, since an analogous cancellation happens for the Pöschl–Teller potential $`V(x)=𝒱\mathrm{cosh}^2(x)`$ (sometimes used as a model for gravitational potentials) for $`\sqrt{\frac{1}{4}𝒱}=\frac{1}{2},\frac{3}{2},\frac{5}{2},\mathrm{}`$ (e.g.,). While the mode situation at $`\mathrm{\Omega }`$ for the RWE alone thus is unclear already, SUSY leaves its relationship to the ZE nontrivial at $`\omega ^2=\mathrm{\Omega }^2`$, $`\xi _1`$ spanning the kernel of the SUSY operator $`d_x+W`$. It should be noted that the existing literature rarely emphasizes this potential difference in spectrum between the two equations: for instance, the captions to Figs. 1 and 2 and Table 1 of merely say ‘Schwarzschild QNMs,’ although the accompanying text makes it clear that in fact the RWE has been studied. In view of this state of affairs, it seems imperative to start from first principles, using comparatively rigorous methods, and study each of the six questions: at $`\mathrm{\Omega }`$, do the RWE and the ZE have a QNM and/or TTM<sub>R</sub> and/or TTM<sub>L</sub>? If this investigation can clarify the pitfalls awaiting a numerical analysis, it will be of value beyond this immediate scope. We will settle the issue as follows: at $`\mathrm{\Omega }`$, the RWE has no modes at all, while the ZE has one mode $`\stackrel{~}{\xi }_1`$ which is simultaneously a QNM and a TTM<sub>L</sub>. To this end, in Section II we first describe some intricacies in the general analysis of (1) if $`V(x)`$ is not finitely supported (as would be natural to assume, e.g., in models of cavity QED), and set the stage by giving the exact solution of the RWE at $`\mathrm{\Omega }`$. The former (Section II A) is necessarily treated in detail, since the subsequent analysis will turn out to involve precisely the most subtle aspects of the theory. However, readers familiar with the issue, or only interested in results on black holes, can initially skip over this part. In Section III, the RWE is studied on the left (near $`r=1`$). It is found that the SUSY generator (5), (7) is not outgoing into the horizon, as one might have been tempted to assume; the consequences for SUSY are discussed. A logical question—which will be answered in the negative—is whether any of the physically relevant solutions encountered in the analysis would be distinguished by its behavior near the black-hole singularity $`r=0`$. Since this involves methods similar to the ones at $`r=1`$, and since some of the results are relevant to Section V (there even exists a logical connection back to Section III, but that can be established only in Section VI), the analysis near $`r=0`$ is presented in Section IV. In Section V, the RWE is studied on the right (near $`r=\mathrm{}`$); combination of the results of Sections IIIV will then prove the above claims. In Section VI, our results are compared and combined with the Leaver-series solution for the outgoing wave to infinity. Section VII discusses the special modes of Kerr holes; for $`a>0`$, these are all found to be TTMs, with both a TTM<sub>L</sub> and a TTM<sub>R</sub> (depending on the sign of the spin) present in the lower half $`\omega `$-plane for each $`(\mathrm{},m,a)`$. Hence, for the TTM<sub>R</sub> this property is not conserved in the Schwarzschild limit (since neither the RWE nor the ZE has one, as has been mentioned above), and the reason for this is elucidated. As a consequence of the analysis, it is found that the Schwarzschild special frequency is a limit point not only for the aforementioned TTMs, but for a multiplet of QNMs as well. Finally, Section VIII contains concluding remarks, discussing both the literature cited above and some remaining questions. ## II Preliminaries ### A Outgoing waves for long-range potentials While it is well known that the tails of the potentials (3), (4) make the determination of their modes very hard numerically unless at least $`|Im\omega ||Re\omega |`$, not all authors have sufficiently emphasized several difficulties of principle which occur on the NIA (so-called zero-modes), as already illustrated by the surprising coincidence QNM=TTM<sub>L</sub> at the end of Section I. Define the outgoing waves to the left ($`f`$) and right ($`g`$) as solutions of (1) obeying $$f(x\mathrm{},\omega )1e^{i\omega x},g(x\mathrm{},\omega )1e^{i\omega x}$$ (8) for $`Im\omega >0`$ and further by analytic continuation; these are the functions figuring in the retarded Green’s function $`\stackrel{~}{G}^\mathrm{R}(x,y;\omega )`$. The continuation conserves the asymptotic forms (8) for those $`\omega `$ for which the functions are finite, as long as $`|\pi /2\mathrm{arg}(\omega )|<3\pi /2`$. A solution is said to be incoming at $`\omega `$ iff it is outgoing at $`\omega `$. The situation, as recently examined by us in in the context of SUSY, now is as follows. First, it could happen that some modes are of higher order, that is, under a generic perturbation they would be split up into several modes rather than merely shifted. However, since it will not occur in the present work this possibility will not be pursued here. Second, in general the outgoing waves will have a branch point at $`\omega =0`$ if the tail is stronger than exponential. Eqs. (2)–(4) show that this happens for $`g`$ but not for $`f`$; now the NIA is the only choice for the branch cut which respects the symmetry $`g(\omega ^{})=g(\omega )^{}`$. On the cut, one defines $`g_{\mathrm{l}(\mathrm{r})}(\omega )=lim_{ϵ0}g(\omega ϵ)`$, and $`\delta g(\omega )g_\mathrm{r}(\omega )g_\mathrm{l}(\omega )`$. Since (8) shows that the increasing parts of $`g_{\mathrm{l}/\mathrm{r}}`$ (if present at all, see below) cancel in $`\delta g`$, and since for each non-real $`\omega `$ the KGE (1) has a unique decreasing solution, one must have proportionality $$\delta g(x,\omega )=\alpha (\omega )g(x,\omega ),$$ (9) with $`\alpha `$ purely imaginary. For each $`\omega ^{}`$ one further defines a discontinuity index $`\mu `$ by $`\delta g(\omega )(\omega \omega ^{})^\mu `$. Generically one expects $`\mu =0`$ while no cut at all would amount to $`\mu =\mathrm{}`$, but at isolated frequencies, corresponding to the zeros of $`\alpha `$, positive integer values for $`\mu `$ are possible as well. For further developments, see Appendix A. Third, if the potentials are not oscillating then frequencies $`\omega ^{}=i|\omega ^{}|`$ are the only candidates for anomalous points, where $$f(\omega )=\frac{𝖿(\omega )}{(\omega \omega ^{})^\nu },$$ (10) with $`𝖿(\omega )`$ finite and nonzero near $`\omega =\omega ^{}`$ and with the anomalous-point index $`\nu 1`$ (of course the same can happen for $`g`$, but this will not occur in the present paper). While these divergences in the outgoing wave are reasonably well known, up to now it apparently has not been realized that they are a mere artifact of the normalization of $`f`$, so that for instance in $`\stackrel{~}{G}^\mathrm{R}(\omega ^{})`$ they would cancel against the corresponding divergence in the normalizing Wronskian. Rather, the true outgoing solution to the left at the anomalous point is $`𝖿(\omega ^{})`$. Since (8) and (10) show that $`𝖿(\omega )(\omega \omega ^{})^\nu e^{i\omega x}`$, it follows that the increasing part of $`𝖿(\omega ^{})`$ vanishes. Hence, $`𝖿(\omega ^{})`$ is actually decreasing for $`x\mathrm{}`$ and thus coincides with the incoming wave: $`𝖿(\omega ^{})f(\omega ^{})`$. Because of this counterintuitive behavior, $`𝖿(\omega ^{})`$ is called anomalous outgoing. It should be noted that $`lim_{\omega \omega ^{}}lim_x\mathrm{}|𝖿(x,\omega )|=\mathrm{}`$ but $`lim_x\mathrm{}lim_{\omega \omega ^{}}|𝖿(x,\omega )|=0`$, severely hampering any brute-force numerical analysis. However, $`𝖿(x,\omega )`$ is completely regular near $`\omega ^{}`$ in any finite region of $`x`$. Physically, the above merely means that at the anomalous point $`\omega ^{}`$, the small (typically exponential) tail scatters so strongly that the outgoing wave is completely different from the free outgoing wave; that is, one has a resonance phenomenon in $`1+1`$ dimensions. Technically, the derivation just given is almost identical to the one of the simple proportionality $`\delta g(x,\omega )g(x,\omega )`$ in the preceding paragraph. However, the concept of anomalous points, while central to this paper, can initially be confusing since the definitions of outgoing and incoming waves near (8) would at first sight seem to be incompatible unless $`\omega =0`$. One can, if necessary, convince oneself by taking $`V(x)=V_0e^{\lambda x}`$ ($`\lambda >0`$) and calculate at least $`k`$ Born approximations to $`f(\omega )`$ to see the anomalous point at $`\omega ^{}=ik\lambda /2`$ ($`k=1`$ being the simplest example). Rescaling $`f𝖿`$ as in the preceding to remove the anomalous-point divergence, one observes that $`𝖿(\omega ^{})`$ and $`f(\omega ^{})`$ agree to any desired order, as they must. Considering the exact solutions available in this or in the Pöschl–Teller case, using standard identities for the Bessel and hypergeometric functions involved one arrives at the same conclusion. These examples also introduce the following approach: if $`V(\lambda ^1\mathrm{ln}z)`$, continued from $`z>0`$, is analytic near $`z=0`$ for certain $`\lambda >0`$ (the assumption of a decreasing potential tail implying that $`V(z=0)=0`$), the KGE (1) has a regular singular point at $`z=0`$ in the variable $`z=e^{\lambda x}`$, and the index equation yields the associated characteristic exponents simply as $`\pm i\omega /\lambda `$ (from now on $`\lambda =1`$ not to burden the notation). The Frobenius theory of regular singular points now distinguishes three cases. (a) In the generic case $`2i\omega 𝐙`$, there are two independent generalized power-series solutions, corresponding to the Born series for the outgoing and incoming waves. If $`2i\omega 𝐙`$, the small ($`z^{i\omega }`$ for $`Im\omega <0`$) solution still is a generalized power series; this is the incoming wave $`f(\omega )`$. However, the large solution is of the form $`_{j=0}^{\mathrm{}}c_jz^{ji\omega }+\zeta f(z,\omega )\mathrm{ln}z`$, where $`\{c_{j1}\}`$ and $`\zeta `$ have to be calculated case by case. (b) Typically $`\zeta 0`$, so that there is no generalized power series for the large solution; this is exactly the case of diverging Born series discussed before, and hence corresponds to an anomalous point. Indeed, in the example of the preceding paragraph these were located at $`\omega ^{}=ik/2`$. The outgoing wave, defined by analytic continuation, will certainly not be logarithmic in $`z`$ and hence should correspond to the small solution, as has been deduced already for anomalous points in general. (c) It can also happen that a ‘miracle’ makes $`\zeta =0`$ even though $`2i\omega 𝐙`$. Hence, in this doubly nongeneric situation, there is a one-parameter family of large generalized power-series solutions, undetermined up to a multiple of the small solution. One of these will correspond to the analytically continued outgoing wave, since in the coordinate $`x`$ the vanishing of $`\zeta `$ means that higher-order corrections to $`V(x)V_0e^x`$ have conspired in a ‘miraculous’ cancellation of the divergence in $`f(\omega )`$. Thus, case (c) does not correspond to an anomalous point. While the Frobenius method in itself of course is elementary, the above recapitulation emphasizes its aspects when $`\omega `$ is not a fixed parameter, but a variable of the theory. Let us complement the discussion by two remarks, one more abstract, the other more computational and concrete. The first remark is that one can analytically continue the solutions along a circle around $`z=0`$, and study the so-called monodromy map this generates in the two-dimensional solution space. For generalized power series the effect follows from the leading coefficient as $`z^{\pm i\omega }(ze^{2\pi i})^{\pm i\omega }=z^{\pm i\omega }e^{2\pi \omega }`$. Hence, in the generic case (a), the monodromy map is a diagonalizable matrix with distinct eigenvalues $`e^{2\pi \omega }`$. On the other hand, in the anomalous-point case (b) this map is a nontrivial Jordan block (e.g.,), because applying it to any large solution yields an image which is not even proportional, as the continuation carries one onto a different branch of the logarithm. The missing eigenvector precisely corresponds to the missing preferred large solution, since both the incoming and outgoing waves are small. This also shows that, if one were to define a preferred ‘interesting’ large solution by other means, such a hypothetical notion of ‘interesting’ cannot be invariant under monodromy, which limits its relevance. Finally, for a ‘miracle’ (c) the monodromy map becomes either $`𝟙`$ (for half-integer $`i\omega `$) or $`+𝟙`$ (for integer $`i\omega `$), so that any solution is an eigenvector. Hence, in the latter case, at the immediate level only the incoming solution is distinguished as the small one, while the preferred increasing solution—the outgoing wave—must be determined by other means. The second remark concerns precisely the calculation of such miraculous outgoing waves. Generically the coefficients of $`f(\omega )=_{j=0}^{\mathrm{}}c_jz^{ji\omega }`$ follow from the recurrence relation $`j(j2i\omega )c_j=_{p=1}^jV_pc_{jp}`$, with $`V(z)=_{p=1}^{\mathrm{}}V_pz^p`$ and $`c_01`$. If $`\omega =\omega ^{}=ik/2`$ ($`k𝐙`$), the brute-force approach is to calculate $`c_k(\omega )`$ from this recurrence for unevaluated $`\omega `$, and finally investigate the limit $`\omega \omega ^{}`$. However, even using computer algebra such operations on (big) rational functions are less transparent than those on (large) integers only. Of course, it is standard to verify whether a miracle occurs by checking if $`_{p=1}^kV_pc_{kp}(\omega ^{})`$ vanishes. If it does, in this way one has verified that a large outgoing wave exists. However, even to determine which value of $`c_k`$ corresponds to $`f(\omega ^{})`$ brute force is not needed. Namely, using de l’Hospital’s rule one finds $`c_k(\omega ^{})=(2ik)^1_{p=1}^kV_p_\omega c_{kp}|_\omega ^{}`$, where for $`1pk1`$ the numbers $`_\omega c_p|_\omega ^{}`$ follow from the $`\omega `$-derivative of the recursion relation, which in the given range of $`p`$ can never lead to any divergences. Finally, the $`c_{jk+1}`$ are calculated from the recursion relation without further ado. For the RWE, in the following sections we will obtain complete analytical control, and a double check (which we have performed) using brute-force computer algebra is still feasible for $`\mathrm{}=2,3`$. However, in a more complicated situation in Section VII B we shall use an approach analogous to the above to calculate (93). When it applies (as it does for black holes), Frobenius theory thus yields complete information on anomalous points and miracles. However, other cases are also possible, such as $`V(x)xe^{\lambda x}`$ (readily handled as $`_\lambda [V(x)e^{\lambda x}]`$, yielding an index $`\nu =2`$), and the responsible exponential tail could even be buried beneath, e.g., an algebraic one; see further Section V A. ### B Exact solution of the Regge–Wheeler equation To avoid the transcendental (2) we write the RWE in terms of $`r`$, $`[r^2(r1)^2d_r^2+r(r1)d_r`$ (11) $`(r1)\{2(n+1)r3\}+\omega ^2r^4]\varphi =0,`$ (12) while (8) translates to $`f(r,\omega )[e(r1)]^{i\omega }`$ and $`g(r,\omega )(re^r)^{i\omega }`$. At $`\omega ^2=\mathrm{\Omega }^2`$, integration of (5)–(7) yields $`\xi _1`$ as the $`\gamma =\gamma _10`$ case in the exact solution $`\xi (r,\gamma )`$ $`=`$ $`[(r1)e^r]^{N/2}{\displaystyle \frac{2nr+3}{r}}\left[1+\gamma I(r)\right],`$ (13) $`I(r)`$ $``$ $`{\displaystyle _1^r}{\displaystyle \frac{dt}{t1}}[(t1)e^t]^N{\displaystyle \frac{t^3}{(2nt+3)^2}},`$ (14) readily found by varying the constant in $`\xi _1`$. The remaining quadrature seems to make $`\xi (r,\gamma )`$ singular near $`2nr+3=0`$; however, in fact this cannot happen since the RWE (12) is regular there, so the singularity must cancel because $`n`$ and $`N`$ in (14) are related as in (6). Motivated by these remarks, one finds that the integral actually is elementary, viz., $`I(r)`$ $`=`$ $`{\displaystyle \frac{(2n1)(n+1)(N2)!}{3nN^{N+1}}}[e^Ne^{Nr}{\displaystyle \underset{j=0}{\overset{N2}{}}}{\displaystyle \frac{[N(1r)]^j}{j!}}]`$ (16) $`+{\displaystyle \frac{e^{Nr}(r1)^{N1}[2nr^2(2n+3)r+6]}{4n^2N(2nr+3)}}.`$ Various choices of the free parameter $`\gamma `$ now will correspond to the various physically interesting solutions. We have already defined $`\xi _1`$ with $`\gamma _1=0`$ as the SUSY generator (7). Since incoming waves are readily and uniquely characterized by being decreasing, their determination is immediate. In particular we see that $`\xi _1`$ itself is incoming from infinity, and from (13) we get in detail $`\xi _1=2ng(\mathrm{\Omega })`$. Similarly, the form (13) shows that $`\xi _2lim_\gamma \mathrm{}\xi (\gamma )/\gamma =f(\mathrm{\Omega })/[(2n+3)N]`$ is incoming from the horizon, i.e., formally $`\gamma _2\mathrm{}`$. Further, we define $`\xi _3`$ as the outgoing wave into the horizon $`f(\mathrm{\Omega })`$, $`\xi _4`$ as the unique small solution near $`r=0`$, and $`\xi _5`$ as the outgoing wave to infinity $`g(\mathrm{\Omega })`$. It may be helpful to note that for $`1j5`$, $`\xi _j`$ is evaluated in Section $`j`$. For general $`\omega `$, the RWE (12) has regular singular points at $`r=1`$ and $`r=0`$, and an irregular singular point at $`r=\mathrm{}`$. We shall successively study these three singularities (in fact the behaviors of the solutions at each of them will turn out to be related); upon combination, the results will completely elucidate the situation at the special frequency and in particular prove the claims made in the Introduction. ## III Near the horizon At the horizon $`r=1`$, one has exactly the situation described in Section II A (with $`\lambda =1`$), so that the index equation has roots $`\pm i\omega `$. Hence, for $`2i\omega =k𝐙\backslash \{0\}`$, typically the large ($`(r1)^{|\omega |}`$) solution is expected to contain a contribution $`(r1)^{|\omega |}\mathrm{ln}(r1)`$ so that $`\omega =ik/2`$ is an anomalous point, as is indeed readily verified for the lowest few $`k`$. However, one now sees at once that, for any $`\mathrm{}`$, $`\omega =\mathrm{\Omega }`$ is a doubly nongeneric, ‘miraculous’ frequency, for the general solution $`\xi (r,\gamma )`$ of (13) contains no log-terms at all. Hence, the Born series is finite and convergent, and the outgoing solution $`f(\mathrm{\Omega })`$ (up to this stage: whatever it may turn out to be) is large so that $`\omega =\mathrm{\Omega }`$ is not an anomalous point. If it seems surprising that, e.g., the simple form of $`\xi _1`$ alone can lead to such a nontrivial conclusion, one can of course for a given $`\mathrm{}`$ evaluate the first $`N1`$ terms of the series for $`f(\mathrm{\Omega })`$ to see the pertinent cancellation (Section II A) occur. In contrast to the incoming wave, in the presence of potential tails the outgoing wave is no longer characterized by its leading asymptotic behavior, nor does the solution (13) at a single frequency allow the continuation procedure stipulated below (8). An elegant indirect method to find the outgoing wave for $`r1`$ is as follows: $`\xi `$ is outgoing iff $`_\omega [(r1)^{i\omega }\xi ]`$ contains no log-terms. Namely, in the general superposition $`_\omega [(r1)^{i\omega }\{Af(\omega )+Bf(\omega )\}]`$ only the second term in square brackets contains $`\omega `$ in an exponent, so that these log-terms come from the incoming part of $`\xi `$. Writing things out using the variation-of-constant solution for $`_\omega \xi `$, the criterion becomes that $`[r/(r1)]\xi ^2(r)`$ (the first factor is a Jacobian) should not contain an $`(r1)^1`$-term. Substituting (13) into this nonlinear function of $`\xi `$, one finds that this happens iff $`\gamma =\gamma _3`$ $``$ $`[{\displaystyle \frac{9}{2}}{\displaystyle \underset{j=0}{\overset{N2}{}}}{\displaystyle \frac{N^{j+1}}{j!}}+3(2n+3){\displaystyle \frac{N^{N+1}}{N!}}]e^N`$ (17) $`=`$ $`{\displaystyle \frac{2811093}{3}}e^8\text{for }\mathrm{}=2;`$ (18) in particular, $`\gamma _3`$ is a sum of negative terms and thus is nonzero for any $`\mathrm{}`$. Hence, defining $`\xi _3\xi (\gamma _3)=(2n+3)f(\mathrm{\Omega })`$, for all $`\mathrm{}`$ one has $`\xi _1\propto ̸\xi _3\propto ̸\xi _2`$, where the former inequality means that the RWE has no TTM<sub>R</sub> at $`\mathrm{\Omega }`$. Summarizing, we are led to the striking conclusion that in spite of its appealing simple form the SUSY generator $`\xi _1`$ is not outgoing into the horizon. The special case (18) for the ‘magic number’ $`\gamma _3`$ can be and has been readily verified by using computer algebra to do an eighth-order Born approximation with unevaluated $`\omega `$, or its refinement described at the end of Section II A. However, for the next-simplest case $`\mathrm{}=3`$ one already needs fortieth order. Despite this computational demand, these series methods are a reliable way of computing $`f(\omega )`$, the logarithmic derivative of which can be compared to the one of $`g(\omega )`$, the latter being stably obtained by integrating the RWE down from large $`r`$. In this way we have verified that for $`\mathrm{}=2`$ there is also no TTM<sub>R</sub> near $`\mathrm{\Omega }`$ on the NIA (say between the nearest two anomalous points, although the rest of the semi-axis is readily checked as well). In fact, for any $`\mathrm{}`$ one readily convinces oneself analytically that the real zero of $`f(x,\mathrm{\Omega })`$ (cf. the asymptotics of $`\xi _3`$ for $`x\pm \mathrm{}`$) is stable under small imaginary perturbations of $`\mathrm{\Omega }`$. Since for a repulsive potential such a zero means that $`f(\omega \mathrm{\Omega })`$ cannot simultaneously by incoming from the right, the claim of on the presence of a TTM<sub>R</sub> is refuted. The above is enough to classify the RWE$``$ZE SUSY-transform on the left. Namely, $`\mathrm{\Omega }`$ has turned out not to be anomalous, while the generator $`\xi _1`$ is mixed (i.e., a nontrivial superposition of incoming and outgoing). This means that the transformation is of category (b2) in the terminology of Appendix B3 in, the result quoted in Section VI of that paper. We then get the prediction that the ‘miracle’ will not repeat itself for the ZE, where on the left $`\mathrm{\Omega }`$ should be an anomalous point with index $`\stackrel{~}{\nu }=1`$, $`\stackrel{~}{\xi }_1`$ being the only solution with no log-terms; besides, it follows from the general theory that $`\stackrel{~}{\xi }_1\stackrel{~}{𝖿}(\mathrm{\Omega })`$ is anomalous outgoing. Independent of the precise situation on the right, this means that for the ZE a TTM<sub>R</sub> at $`\mathrm{\Omega }`$ would simultaneously be an NM, which cannot occur. Hence, also the ZE has no TTM<sub>R</sub> at $`\mathrm{\Omega }`$. Let us conclude this section by pointing out two interesting perspectives on SUSY offered by the preceding. In the first place, in we obtained a very straightforward proof of $`\xi _1`$ not being outgoing into the horizon—a property referring to the RWE only—by looking at the anomalous point of the ZE. In the second place, in Section II A it has been argued that this very anomalous point is a nuisance numerically near $`\mathrm{\Omega }`$. However, in this section we have seen that the corresponding anomalous point is absent in the RWE, and any numerical result obtained there can subsequently be SUSY-transformed back to the ZE. Thus, SUSY can be used both as an analytical tool for the RWE and as a numerical tool for the ZE. The theory of shows that the latter application actually has some generality, namely, as long as the anomalous-outgoing wave is nodeless (as it has to be for repulsive $`\stackrel{~}{V}`$, cf.). ## IV Near the black-hole singularity The index equation now gives the characteristic exponents $`\rho _1=3`$, $`\rho _2=1`$, so that $`\rho _1\rho _2𝐙`$. While this is still widely recognized (e.g.,), previous works do not seem to carry out the associated series expansion for a more precise determination of the character of the singularity, i.e., logarithmic/anomalous (case (b) of Section II A) or miraculous (case (c)). The above shows that this determination in fact is decisively simpler than for $`r=1`$, since the expansion is only needed to fourth order for any $`\mathrm{}`$ and $`\omega `$. By again merely looking at the form of (13) (cf. Section III), one sees that at least for $`\omega =\mathrm{\Omega }`$ a miracle happens at $`r=0`$, since the general solution then contains no $`\mathrm{ln}r`$-terms. Indeed, actually carrying out the fourth-order expansion readily shows that a power-series solution $`r^1`$ exists if and only if $`\omega ^2=\mathrm{\Omega }^2`$. Let us now define $`\xi _4`$ as the unique solution which is finite near $`r=0`$: $`\xi _4(r)`$ $`=`$ $`\left({\displaystyle \frac{e^r}{r1}}\right)^{N/2}{\displaystyle \frac{2nr+3}{r}}{\displaystyle _0^r}{\displaystyle \frac{dt}{t1}}{\displaystyle \frac{[(t1)e^t]^Nt^3}{(2nt+3)^2}}`$ (19) $`=`$ $`\xi _1(r)[I(0)+I(r)]`$ (20) $`=`$ $`I(0)\xi _1(r)+\xi _2(r),`$ (21) so that $`\gamma _4I(0)^1`$. Obviously $`\xi _4\propto ̸\xi _1`$, since their $`r\mathrm{}`$ asymptotics differ. Also $`\xi _4\propto ̸\xi _2`$ since $`I(0)0`$, all factors of the integrand having a fixed sign on $`(0,1)`$. Could $`\xi _4\xi _3=\xi _1+\gamma _3\xi _2`$, that is, could the unexpected result $`\xi _3\xi _1`$ for the outgoing wave into the horizon perhaps be distinguished by its behavior near the black-hole singularity? This is true iff $`I(0)=\gamma _3^1`$. We will now prove that this cannot happen; for the required bound, (16) is less convenient than direct estimation of the integral. Because $`N`$ is even one has $`I(0)>0`$, and $`I(0)<\frac{1}{9}_0^1𝑑t(1t)^{N1}e^{Nt}t^3`$; use $`_0^1𝑑t(1t)^ke^{Nt}=(k!/N^{k+1})[e^N_{j=0}^kN^j/j!]`$, doing the terms with $`NjN+2`$ exactly. In the remainder use $`e^N_{j=0}^{N1}N^j/j!>0`$, then the contribution of this remainder is seen to be negative, so it can be omitted in the upper bound. After this, the inequality becomes $`I(0)<2(N+1)/9N^3<1/3N^2`$. For $`|\gamma _3|`$ use $`_{j=0}^{N2}N^j/j!<e^N`$, $`2n+3<3\sqrt{N}`$, and $`N!>2N^{N+1/2}e^N`$, yielding $$|\gamma _3|<9N.$$ (22) Combination now gives $`I(0)|\gamma _3|<3/N<1`$, for $`N8`$. Hence, $`\xi _1`$ through $`\xi _4`$ are all different. ## V Near $`r=\mathrm{}`$; the mode situation ### A A continuation theorem Let us first derive a general result on the KGE (1): if $`V(x)`$ is analytic for $`x>x_0`$ (say), and if its continuation from those $`x`$ has $`V(x)=O(x^{1\eta })`$ for $`|\mathrm{arg}x|<\beta `$ and some $`\beta ,\eta >0`$, then $`g(x,\omega )1e^{i\omega x}`$ if (A) $`|\mathrm{arg}x|<\beta `$, (B) $`|\pi /2\mathrm{arg}\omega |<\pi /2+\beta `$, and (C) $`|\pi /2\mathrm{arg}\omega x|<3\pi /2`$. In particular, $`g(\omega )`$ has no anomalous points for $`|\pi /2\mathrm{arg}\omega |<\mathrm{min}(\pi /2+\beta ,3\pi /2)`$. (The counterpart for $`f`$ is obvious but not needed here.) Namely, for $`\omega `$ obeying (B) the continuation of (8) can be first carried out in parts of the $`x`$-plane where it merely amounts to selecting and normalizing the unique small solution, which can never lead to any divergence. This proves the result for $`|\pi /2\mathrm{arg}\omega x|<\pi /2`$, which in a second step can be extended at fixed $`\omega `$ by trivially solving (1) in a region where $`V`$ is negligible; condition (C) (implied by (A) and (B) if $`\beta \pi /2`$) arises because the asymptotic form for $`g`$ cannot be continued in this way from a region where it dominates across an anti-Stokes line. In applications, the second step would typically take the form of continuation back to the physical region $`x>x_0`$. The rationale for the two-step procedure is seen to be that continuations in $`x`$ are easier than those in $`\omega `$, since for the former one has the linear ordinary differential equation (1) at one’s disposal. For an example, calculating a few terms in the Born series shows the theorem in action for $`V(x>x_0)=e^{\lambda x}\mathrm{cos}(\mu x)`$, which has $`\beta =\mathrm{arctan}(\lambda /|\mu |)`$. ### B Application to the Regge–Wheeler equation Turning to the RWE, the above procedure is best carried out in terms of $`r`$. Since $`V(r)`$ vanishes if $`|r|\mathrm{}`$, one has $`\beta =\mathrm{}`$ so that $`g`$ has no anomalous points, its asymptotics being as below (12) for $`|\pi /2\mathrm{arg}\omega r|<3\pi /2`$. This should now be compared with the asymptotics of $`\xi `$, for which (13) and (16) show that $`\xi (r,\gamma )`$ $``$ $`2n\left[1+\gamma {\displaystyle \frac{(2n1)(n+1)(N2)!e^N}{3nN^{N+1}}}\right](re^r)^{N/2}`$ (24) $`+{\displaystyle \frac{\gamma }{2nN}}(re^r)^{N/2}`$ in the entire $`r`$-plane. In particular, for both $`g_\mathrm{l}`$ (which has $`\mathrm{arg}\mathrm{\Omega }=3\pi /2`$) and $`g_\mathrm{r}`$ (with $`\mathrm{arg}\mathrm{\Omega }=\pi /2`$) one can make the comparison with (24) in the region $`Rer0`$, where the asymptotics are not dominated by the second term on the r.h.s., as well as for the original $`x>0`$. This of course reproduces that $`\xi (\gamma )`$ is incoming from infinity for $`\gamma =0`$, but also shows that it is outgoing for both branches of $`g(\omega )`$ iff $$\gamma =\gamma _5\frac{3nN^{N+1}e^N}{(12n)(n+1)(N2)!},$$ (25) in which case $`\xi _5\xi (\gamma _5)=(\gamma _5/2nN)g_{\mathrm{l}/\mathrm{r}}(\mathrm{\Omega })`$, in accordance with the general statement (see, Appendix B2) that if $`g_\mathrm{l}`$ and $`g_\mathrm{r}`$ coincide they must be real. Hence, the discontinuity index $`\mu 1`$ at $`\mathrm{\Omega }`$. This statement on $`\mu `$ (dealing with analytic continuations in $`\omega `$ and with $`r\mathrm{}`$) has thus been related to the ‘miracles’ near $`r=1`$ and $`r=0`$ (dealing with continuations for small $`r`$ at a fixed $`\omega `$), since the latter make that $`\xi (r,\gamma )`$ has the same asymptotics for $`\mathrm{arg}r<\pi /2`$ and $`\mathrm{arg}r>\pi /2`$. Note that for a qualitatively similar but otherwise arbitrary potential, the miracle is a ‘leftist’ and the vanishing of $`\alpha (\omega )`$ a ‘rightist’ concept, so that the two would be completely unrelated. Of course, the RW potential is anything but arbitrary, being of the form $`V(x)=W^2(x)d_xW(x)+\mathrm{\Omega }^2`$, as follows at once from (2), (3), and (5) or indeed from SUSY in general. Obviously $`\xi _1\propto ̸\xi _5\propto ̸\xi _2`$, where the former inequality means that $`\mathrm{\Omega }`$ indeed is not anomalous for the RWE on the right, and the latter that the RWE has no TTM<sub>L</sub> at $`\mathrm{\Omega }`$. Two questions remain, namely, whether possibly $`\gamma _5=\gamma _3`$ or $`\gamma _5=\gamma _4`$; the answer is negative both times. For the latter (Section IV showed that the outgoing wave into the horizon was not distinguished near $`r=0`$, but could perhaps the outgoing wave to infinity be?), this is immediate since (16) for $`I(0)=\gamma _4^1`$ shows that $`\gamma _4^1\gamma _5^1`$ is a sum of positive terms, hence nonzero. For the former, using methods as above (22) (in particular, $`N!<\sqrt{8}N^{N+1/2}e^N`$) leads to $`|\gamma _5|>(\sqrt{3}/4)N^2`$. If $`\mathrm{}3N40`$, this implies $`|\gamma _5|>9N>|\gamma _3|`$ (cf. (22)), while for the remaining case $`\mathrm{}=2`$ the inequality $`|\gamma _5|=e^82^{24}/135>|\gamma _3|`$ (cf. (18)) is immediate. Hence, $`\xi _5\propto ̸\xi _3`$ and the RWE has no QNM at $`\mathrm{\Omega }`$, even though the positive discontinuity index $`\mu 1`$ in itself would have allowed this. Thus, the five solutions we have identified—while of course not independent—are all different, and the surprise is rather that there are not six of them, i.e., that $`g_\mathrm{l}(\mathrm{\Omega })=g_\mathrm{r}(\mathrm{\Omega })`$, given that the outgoing wave for the Schwarzschild black hole definitely has a branch cut in general (see and Section V D). ### C SUSY and the Zerilli equation The mere absence of anomalous points in $`g`$ for the RWE (cf. above (24)) already implies that the SUSY generated by $`\xi _1`$ is of type (a1) on the right in the terminology of. The general theory then shows that $`\stackrel{~}{\xi }_1\stackrel{~}{g}_\mathrm{l}(\mathrm{\Omega })=\stackrel{~}{g}_\mathrm{r}(\mathrm{\Omega })`$. Since we have shown in Section III that it is also anomalous outgoing on the left, $`\stackrel{~}{\xi }_1`$ thus is a QNM=TTM<sub>L</sub> of the ZE (of course, this once more shows that the ZE has no TTM<sub>R</sub> at $`\mathrm{\Omega }`$). Because $`\mu 1`$ already in the original system, the equality of $`\stackrel{~}{g}_{\mathrm{l}/\mathrm{r}}`$ in the above can immediately be sharpened to $`\stackrel{~}{\mu }2`$ by the elementary SUSY relationship $$\stackrel{~}{\alpha }(\omega )=\frac{\mathrm{\Omega }\omega }{\mathrm{\Omega }+\omega }\alpha (\omega ).$$ (26) However, to determine the orders of the modes one needs the full apparatus of. Its Eqs. (B6), (B7) show that the QNM Wronskian $`\stackrel{~}{J}_\mathrm{q}(\omega )`$ is regular near $`\mathrm{\Omega }`$, while the TTM Wronskian $`\stackrel{~}{J}_\mathrm{t}(\omega )`$ has a first-order zero there and a simple pole at $`\mathrm{\Omega }`$. Accounting for the anomalous point of $`\stackrel{~}{f}`$, this means that both the QNM and the TTM<sub>L</sub> of the ZE furnished by $`\stackrel{~}{\xi }_1`$ are simple—intuitively plausible, since the RWE has no QNMs/TTM<sub>L</sub>s to double the ones generated by SUSY. The proportionalities involved are elucidated as $`\stackrel{~}{\xi }_1=\stackrel{~}{\xi }_2=\stackrel{~}{\xi }_3/\gamma _3=\stackrel{~}{\xi }_4=\stackrel{~}{\xi }_5/\gamma _5=\stackrel{~}{f}(\mathrm{\Omega })/(2n+3)=i(2n+3)\gamma _3^1\stackrel{~}{𝖿}(\mathrm{\Omega })=\stackrel{~}{g}(\mathrm{\Omega })/(2n)`$. Incidentally, since $`\stackrel{~}{V}(|r|\mathrm{})`$ vanishes just as well as the RW potential, $`\stackrel{~}{g}`$ has the same asymptotics as established above (24) for $`g`$. Analogously continuing $`\stackrel{~}{\xi }_1(r)`$ to $`Rer0`$ then reproduces that this function is outgoing on the right, without having to calculate the general solution of the ZE at $`\mathrm{\Omega }^2`$ and independent of SUSY. This also shows why the ZE having $`\stackrel{~}{\alpha }(\mathrm{\Omega })=0`$ in the absence of a ‘miracle’ does not contradict the discussion below (25), relating the two. Namely, in this case $`\stackrel{~}{g}`$, being a TTM<sub>L</sub>, is small and hence single-valued near the branch point of the general solution. For readers mainly interested in results, it should be pointed out that this subsection inter alia proves that the ZE retarded Green’s function has a pole at $`\omega =\mathrm{\Omega }`$, the statement of which is completely independent of the intricacies of anomalous outgoing waves. ### D Discontinuity indices The above has established the mode situation at the special frequency. It will now be shown that the derived inequalities for the discontinuity indices are in fact realized as equalities, i.e., that one has $`\mu =1`$, which implies $`\stackrel{~}{\mu }=2`$ by (26). To this end, define $`g_1(\omega )_\omega g(\omega )`$ and make the variation-of-constant Ansatz $`g_1(\mathrm{\Omega })=g(\mathrm{\Omega })h(\mathrm{\Omega })`$. One finds $$d_xh(x,\mathrm{\Omega })=2\mathrm{\Omega }g^2(r,\mathrm{\Omega })_{r_0}^r𝑑t\frac{t}{t1}g^2(t,\mathrm{\Omega }),$$ (27) and to get the correct asymptotics $`d_xh(x,\mathrm{\Omega })i`$ also if $`r`$ is continued to the left half-plane, one should set $`r_0=\mathrm{}`$. However, for $`g_{1\mathrm{r}}(r,\mathrm{\Omega })`$ ($`g_{1\mathrm{l}}(r,\mathrm{\Omega })`$) the asymptotics are imposed if $`r`$ is continued into the second (third) quadrant, so that for $`r>0`$ one must put $`tt+iϵ`$ ($`ttiϵ`$) in the integrand of (27). Evaluating the residues at $`t=1`$ and $`t=0`$, one finds in detail $$\delta _{\mathrm{}}^{r>1}𝑑t\frac{t\xi _5^2(t)}{t1}=2\pi i\left[\frac{2}{N}(\gamma _3\gamma _5)+9\left(\frac{\gamma _5}{\gamma _4}1\right)^2\right]$$ (28) (see Section II A for the definition of $`\delta `$). In particular, since it has already been shown in Section V B that $`|\gamma _5|>|\gamma _3|`$, the first term in brackets of (28) (which is due to the pole at $`t=1`$, and hence anticipated to be $`\gamma _3\gamma _5`$) is positive as well as the second so that $`\delta g_1(\mathrm{\Omega })0`$ and $`\mu =1`$. Eqs. (27) and (28) will have an interesting application in Section VIII. ## VI Leaver series In we already touched upon the relation of this work to the first Leaver series $$f(r,\omega )\left(\frac{r^2e^r}{r1}\right)^{i\omega }\underset{j=0}{\overset{\mathrm{}}{}}c_j\left(\frac{r1}{r}\right)^j.$$ (29) However, at least as interesting a connection exists with the second Leaver series, $`g(r,\omega )`$ $``$ $`r^3\left({\displaystyle \frac{e^r}{r1}}\right)^{i\omega }{\displaystyle \underset{j=0}{\overset{\mathrm{}}{}}}c_j(12i\omega )_j`$ (31) $`\times U(3+j2i\omega ,5,2i\omega r)`$ in terms of the same $`c_j`$ as in (29), and one of the more ‘irregular’ cases (integer second argument) of the irregular confluent hypergeometric function, $`U(`$ $`\kappa ,5,z)={\displaystyle \frac{1}{24\mathrm{\Gamma }(\kappa 4)}}[M(\kappa ,5,z)\mathrm{ln}z`$ (34) $`+{\displaystyle \underset{p=0}{\overset{\mathrm{}}{}}}{\displaystyle \frac{(\kappa )_p}{(5)_p}}{\displaystyle \frac{z^p}{p!}}\{\psi (\kappa +p)\psi (1+p)\psi (5+p)\}]`$ $`+{\displaystyle \frac{6}{\mathrm{\Gamma }(\kappa )z^4}}M_4(\kappa 4,3,z),`$ where $`M_4`$ is the series to four terms of the regular confluent hypergeometric function $`M`$, and $`\psi =\mathrm{\Gamma }^{}/\mathrm{\Gamma }`$. To see the connection, let $`2i\omega =L=1,2,\mathrm{}`$, so that the Pochhammer symbol reduces to $`(1L)_j=()^j(L1)!/(Lj1)!`$ for $`0jL1`$, and zero for $`jL`$. Typically this zero is compensated by the anomalous-point divergence in $`c_j`$, so that the series still involves $`U(5,5,z)=e^z\mathrm{\Gamma }(4,z)`$ and higher, which have logarithmic branch points at $`z=0`$ (cf. (34)). However, if $`\omega =\omega _\mathrm{m}`$ is a ‘miraculous’ frequency, the sum in (31) truncates: $`g(r,\omega _\mathrm{m})`$ $``$ $`r^3\left({\displaystyle \frac{e^r}{r1}}\right)^{L/2}{\displaystyle \underset{j=0}{\overset{L1}{}}}c_j{\displaystyle \frac{()^j}{(Lj1)!}}`$ (36) $`\times U(3+jL,5,Lr),`$ in which only those $`c_j`$’s enter which are the same for all solutions which are large near $`r=1`$, so that in particular for $`L=N`$ one has $`c_j=(1/j!)d_u^j|_{u=0}\mathrm{exp}\{N/(u1)\}[2n(1u)^N+3(1u)^{N+1}]`$ (combine (13) and (29)). The functions $`U`$ involved in (36) are evaluated from (34) as $`U(2,5,z)`$ $`=`$ $`z^2+4z^3+6z^4,`$ (37) $`U(1,5,z)`$ $`=`$ $`z^1+3z^2+6z^3+6z^4,`$ (38) $`U(k,5,z)`$ $`=`$ $`{\displaystyle \underset{p=0}{\overset{k}{}}}\left(\begin{array}{c}k\\ p\end{array}\right)(p+5)_{kp}()^{k+p}z^p`$ (42) $`(k=0,1,\mathrm{}),`$ Since $`U(2,5,z)`$ and $`U(1,5,z)`$ are linearly independent and since $`c_{L1}`$ and $`c_{L2}`$ cannot both vanish for a nonzero sequence $`\{c_j\}`$ because of the recursion $`(j+1)(j+1L)c_{j+1}`$ (43) $`[2j^2+(24L)j+2L(L1)+2(n+1)3]c_j`$ (44) $`+(j^22Lj+L^24)c_{j1}=0,`$ (45) one concludes that $`g(r,\omega _\mathrm{m})`$ is a large single-valued solution near $`r=0`$. In Section IV it has already been established that this can only happen for $`\omega _\mathrm{m}^2=\mathrm{\Omega }^2`$, so that $`\mathrm{\Omega }`$ is the only ‘miraculous’ frequency of the RWE. It should be emphasized that this is a nontrivial statement on the singularity near $`r=1`$, and that the above proof involved an analysis both near $`r=0`$ and $`r=\mathrm{}`$. Thus, $`L=N`$ is the only case which actually occurs, and hence the r.h.s. of (36) should evaluate to $`2(n+1)N^3\xi _5(r)`$ as given by (13), (16) and (25), where the constant of proportionality (for the $`c_j`$ below (36)) has been determined from the leading asymptotics for $`r\mathrm{}`$. Verification amounts to comparing two polynomials of degree $`N+1`$ with rational coefficients, which has been done explicitly for $`\mathrm{}=2`$. For general $`\mathrm{}`$, the stated outcome of the sum in (36) can be proved by downward induction in the power of $`r`$, using the recursion (45) upward; however, we omit the laborious details. The expression (31), (34) is also of interest for nonmiraculous $`\omega =i|\omega |`$. Namely, it shows that in the Leaver series for $`\delta g`$ one has $$\delta U(3+j2i\omega ,5,2i\omega r)=\frac{i\pi M(3+j2i\omega ,5,2i\omega r)}{12\mathrm{\Gamma }(j2i\omega 1)}.$$ (46) The relevance of this latter relation is not so much that the regular $`M`$ is easier to handle than the $`U`$ function (which it is), but rather that it points to a computational scheme for $`\delta g`$ directly and not as an exponentially small difference between the $`g_{\mathrm{l}/\mathrm{r}}`$. In particular, the proportionality derived in Section II A means that one can evaluate $`\alpha (\omega )`$ at any convenient $`r>1`$, and a method of calculation which avoids substracting nearly equal large numbers would enable one to look for, e.g., possible simplifications as $`r\mathrm{}`$. However, the properties of the (non-truncating) sum (31) itself on the NIA are still under investigation. In its turn, this last remark also motivates the explicit verification of (36) in the previous paragraph. ## VII Kerr holes ### A Kerr-hole special modes Both the exact solution for the stationary metric and the separated wave equations for its linearized perturbations can be generalized from the spherically symmetric Schwarzschild hole to the axisymmetric, rotating Kerr hole. Presently we investigate the issues addressed in the preceding sections in this wider context, and start by establishing some notation. Expanding the pertinent effective scalar field as $$\psi (\omega ,r,\theta ,\varphi )=\underset{\mathrm{}=2}{\overset{\mathrm{}}{}}\underset{m=\mathrm{}}{\overset{\mathrm{}}{}}e^{im\varphi }S(\theta ,\omega )R(r,\omega )$$ (47) in Boyer–Lindquist coordinates, the radial equation reads $`\mathrm{\Delta }^2d_r\left({\displaystyle \frac{d_rP}{\mathrm{\Delta }}}\right)+[{\displaystyle \frac{K^2+is(12r)K}{\mathrm{\Delta }}}+4is\omega r\mathrm{¯}\lambda ]P`$ $`=`$ (48) $`[\mathrm{\Delta }(d_ri{\displaystyle \frac{sK}{2\mathrm{\Delta }}}+{\displaystyle \frac{12r}{\mathrm{\Delta }}})\left(d_r+i{\displaystyle \frac{sK}{2\mathrm{\Delta }}}\right)+3is\omega r\mathrm{¯}\lambda ]P`$ $`=`$ $`0,`$ (49) (50) where we introduced the auxiliary quantities $`\mathrm{\Delta }`$ $`=`$ $`r^2r+a^2`$ (51) $`=`$ $`(rr_+)(rr_{}),`$ (52) $`r_\pm `$ $`=`$ $`{\displaystyle \frac{1\pm b}{2}},`$ (53) $`b`$ $`=`$ $`\sqrt{14a^2},`$ (54) $`K`$ $`=`$ $`(r^2+a^2)\omega am,`$ (55) and where $`P=\mathrm{\Delta }^{1+s/2}R`$. In (50), $`0a<\frac{1}{2}`$ is the black-hole rotation, and $`s=\pm 2`$ is the spin of the field. The radial field $`P`$ depends on $`(r,\omega ,\mathrm{},m,a)`$ and on the sign of $`s`$, but this will only be indicated where necessary. The angular separation constant $`\mathrm{¯}\lambda `$ is an eigenvalue of the equation for $`S(\theta ,\omega )`$, cf. (90) below, and different eigenvalues correspond to different angular-momentum sectors. Since $`\mathrm{¯}\lambda `$ obeys the symmetries $`\mathrm{¯}\lambda _m^{}(\omega )=\mathrm{¯}\lambda _m^{}(\omega ^{})`$, $`\mathrm{¯}\lambda _m(\omega )=\mathrm{¯}\lambda _m(\omega )`$, and $`\mathrm{¯}\lambda _s=\mathrm{¯}\lambda _s`$ (besides, $`\mathrm{¯}\lambda (\omega ,a)=\mathrm{¯}\lambda (\omega a)`$), from (50) it follows that $`P_m^{}(\omega )=P_m^{}(\omega ^{})`$ and $`P_{ms}(\omega )=P_{m,s}(\omega )`$; an advantage of using $`P`$ instead of $`R`$ is that this makes the latter relation manifest. Moreover, there is the considerably deeper symmetry between $`P_s`$ and $`P_s`$ furnished by the Teukolsky–Starobinsky identities (which follow by combining the finite-$`a`$ generalization of (63) below with its inverse for opposite $`s`$, see also). Since the special modes (59) below are in the kernel of the corresponding symmetry operator (cf. (VII B)) we do not give the formulas; however, the existence of this symmetry will be used at the end of Section VII C. Where the mere existence of the ordinary differential equation (50) is already remarkable, this holds a fortiori for its exact solvability if $`\omega =\mathrm{\Omega }(a)`$ is a zero of the so-called Starobinsky constant $`𝒮`$ $`=`$ $`\mathrm{¯}\lambda ^2(\mathrm{¯}\lambda +2)^2+8\mathrm{¯}\lambda [5\mathrm{¯}\lambda (am\omega a^2\omega ^2)+6(am\omega +a^2\omega ^2)]`$ (57) $`+36\omega ^2+144(am\omega a^2\omega ^2)^2.`$ Namely, in that case the special modes $`P=𝒫`$ are given by $`𝒫`$ $`=`$ $`[is\mathrm{\Omega }r^3+\mathrm{¯}\lambda r^2i{\displaystyle \frac{sQ}{8\mathrm{\Omega }}}r2a^2+\frac{2}{3}\mathrm{¯}\lambda (a^2{\displaystyle \frac{am}{\mathrm{\Omega }}})`$ (59) $`{\displaystyle \frac{\mathrm{¯}\lambda +2}{24\mathrm{\Omega }^2}}Q]\left({\displaystyle \frac{rr_+}{rr_{}}}\right)^{i\frac{sam}{2b}}e^{is\mathrm{\Omega }x/2},`$ $`Q`$ $`=`$ $`\mathrm{¯}\lambda (\mathrm{¯}\lambda +2)+12(am\mathrm{\Omega }a^2\mathrm{\Omega }^2)+3is\mathrm{\Omega },`$ (60) with the tortoise coordinate $$x=r+\frac{r_+}{b}\mathrm{ln}(rr_+)\frac{r_{}}{b}\mathrm{ln}(rr_{})$$ (61) mapping $`r_+<r<\mathrm{}`$ to $`x𝐑`$. In view of the aforementioned symmetries, one can take $`Im\mathrm{\Omega }<0`$ for consistency with the preceding sections. When assessing the physical significance of the solutions (59), in comparison with the RWE/ZE case one encounters two technical subtleties, which are readily dealt with. Writing (50) as a KGE in the $`x`$-coordinate for $`\sqrt{r^2+a^2}P/\mathrm{\Delta }`$, one finds that the ensuing potential is (i) complex and frequency dependent, and (ii) tending to a nonzero constant if $`x\mathrm{}`$, and having an $`O(x^1)`$ tail for $`x\mathrm{}`$. (i) If $`V=V(x,\omega )`$, besides the outgoing waves (8) one also has to define incoming waves $`F,G`$ (temporary notation), being solutions of (1) at $`\omega `$ with $`G(x\mathrm{},\omega )e^{i\omega x}`$ for $`Im\omega <0`$ and continued analytically from there, and analogously for $`F`$. In the frequency-independent case these become simply $`G(\omega )=g(\omega )`$ etc., but in general $`G`$ and $`g`$ are unrelated. Fortunately, for (50) one has $`G_{ms}(\omega )=g_{m,s}(\omega )`$. (ii) A potential having a nonzero spatial limit can be handled by a shift in frequency; in particular, for (50) it is customary to define $`\stackrel{~}{\omega }=\omega ma/r_+`$. The $`O(x^1)`$ potential tails give rise to a power-law prefactor in the wave-function asymptotics, not unlike the one occurring for the RWE when $`g`$ is considered as a function of $`r`$, cf. below (12); these power laws will cause no problems here. The character of the special modes on the right is now easily determined. With our convention on $`\mathrm{\Omega }`$, it follows from (59) that $`𝒫_2𝒫_{s=2}`$ is decaying and hence incoming from $`x=+\mathrm{}`$, while $`𝒫_2`$ is outgoing by the continuation argument of Section V A. Near the horizon $`r=r_+`$, the characteristic exponents of (50) read $`\rho _{1(2)}=1\pm (1+isr_+\stackrel{~}{\omega }/2b)`$. For the special modes one thus has $`\rho _1\rho _2=2+isr_+\stackrel{~}{\mathrm{\Omega }}/b`$, with $`\mathrm{\Omega }`$ a zero of (57). For $`a0`$ this becomes $`\rho _1\rho _22+sN/2𝐙`$, but one does not expect the difference to keep this value for nonzero $`a`$. Indeed, the expansion (96) below shows that $`\rho _1\rho _2`$ is non-integer for all sufficiently small nonzero $`a`$. While we have not attempted a proof, it seems very unlikely that, in some $`(\mathrm{},m)`$-sector, the complex function $`\rho _1(a)\rho _2(a)`$ would assume a real integer value exactly on $`0<a<\frac{1}{2}`$. Thus, for nonzero rotation, the algebraically special frequency $`\mathrm{\Omega }(a)`$ is actually generic in the terminology of Section II A. Since (59), (61) show that the modes $`𝒫`$ are generalized power series corresponding to $`\rho _2`$ (equivalently, they are eigenfunctions of the monodromy map near $`r_+`$ with eigenvalue $`e^{2\pi i\rho _2}`$), this means that $`𝒫_2`$ is outgoing into the horizon and hence a TTM<sub>R</sub> for $`a>0`$. Similarly, $`𝒫_2`$ is incoming from the horizon and thus a TTM<sub>L</sub>; this follows already from its being decreasing on the left, so that for $`𝒫_2`$ the condition $`a>0`$ is not needed. For $`a=0`$, in which case the above analysis is inconclusive for $`𝒫_2`$, the radial equation (50) does not reduce to the RWE or ZE, but to the Bardeen–Press equation instead; the explicit mapping between them reads $`P`$ $`=`$ $`r^3[\{(W{\displaystyle \frac{N}{2}})(2nr+2\pm 1){\displaystyle \frac{1}{r}}is\omega \}`$ (63) $`\times \{d_x{\displaystyle \frac{is\omega }{2}}\}+W^2d_xW{\displaystyle \frac{N^2}{4}}]\varphi `$ $``$ $`A_\pm \varphi ,`$ (64) where the upper (lower) signs should be chosen if $`\varphi (r,\omega )`$ is a solution of the RWE (ZE). Calculating the Born series for $`f`$ for (50) with $`\mathrm{}=2`$ (so that $`\mathrm{¯}\lambda =4`$, cf. (92) below), at $`\mathrm{\Omega }=4i`$ one finds a divergence (like for the ZE, but here in sixth order) if $`s=2`$ but not if $`s=+2`$ (like for the RWE, but here the relevant cancellation occurs in tenth order), so that $`\mathrm{\Omega }`$ is anomalous (miraculous) for $`s=2`$ ($`+2`$). However, $`𝒫_2`$ is a nontrivial superposition of outgoing and incoming waves—again involving the ‘magic number’ 11093, cf. (18)—in marked contrast to the situation for any nonzero $`a`$. Thus, the unexpected findings of Section III are not an artifact of the mapping (63)—a kind of $`\omega `$-dependent SUSY transform—but intrinsic to the Schwarzschild limit, which apparently has to be highly singular. To investigate this further, one should study (50)–(59) for small $`a`$; this is done in the next subsection. ### B Schwarzschild limit It should be explained why the $`s=2`$ special mode behaves differently for $`a=0`$ than for all other $`a`$. This is not a consequence of a singularity in (59) itself, which in the limit tends smoothly to $`𝒫_2^{(0)}(r)`$ $`=`$ $`{\displaystyle \frac{1}{6n}}[8n^2(n+1)r^3+12n^2r^2+18nr+9]`$ (67) $`\times e^{Nx/2},`$ $`𝒫_2^{(0)}(r)`$ $`=`$ $`{\displaystyle \frac{2n}{3}}[2(n+1)r^33r^2]e^{Nx/2},`$ (68) where one should note the similarity with (4) and (3) respectively, and where the quantitative relation to the special solutions of the RWE/ZE reads $`A_+\xi _1`$ $`=`$ $`2nN𝒫_2^{(0)}\delta _{s2},`$ (70) $`A_{}\stackrel{~}{\xi }_1`$ $`=`$ $`{\displaystyle \frac{N}{2n}}𝒫_2^{(0)}\delta _{s,2}.`$ (71) Since it has been shown in Section VII A that $`𝒫_2`$ is outgoing on the left and that $`\mathrm{\Omega }(a)`$ is generic for nonzero $`a`$, this implies $`\underset{a0}{lim}\underset{\omega \mathrm{\Omega }(a)}{lim}f_2(\omega ,a)`$ $`=`$ $`{\displaystyle \frac{2n𝒫_2^{(0)}}{(2n+3)(N+1)}}`$ (72) $`=`$ $`{\displaystyle \frac{A_+\xi _1}{(2n+3)N(N+1)}},`$ (73) where $`f_2f_{s=2}`$ has been normalized as $$f_2(x\mathrm{},\omega ,a)1e^{i\stackrel{~}{\omega }x},$$ (74) and where the constant of proportionality has been obtained by combining this last relation with (67) and (70). On the other hand, by the very definition of an outgoing wave, taking the limits in the opposite order yields an outgoing result, i.e., $`\underset{\omega \mathrm{\Omega }}{lim}\underset{a0}{lim}f_2(\omega ,a)`$ $`=`$ $`f_2(\mathrm{\Omega },a=0)`$ (75) $`=`$ $`{\displaystyle \frac{A_+\xi _3}{(2n+3)N(N+1)}},`$ (76) where the constant of proportionality is the same as in (73) because $`\xi _1`$ and $`\xi _3`$ have equal increasing parts, cf. (13). To see how these noncommuting limits arise, set $`f_2(r,\omega )=e^{i\stackrel{~}{\omega }r_+}b^{i\stackrel{~}{\omega }r_{}/b}_{j=0}^{\mathrm{}}c_j(rr_+)^{ji\stackrel{~}{\omega }r_+/b}`$ (i.e., $`c_0=1`$), upon which the $`c_j`$ follow form the recursion $$\alpha _jc_j+\beta _jc_{j1}+\gamma _jc_{j2}+\delta _jc_{j3}+ϵ_jc_{j4}=0,$$ (77) $`\alpha _j`$ $`=`$ $`b^2j\left(j22i\stackrel{~}{\omega }{\displaystyle \frac{r_+}{b}}\right),`$ (78) $`\beta _j`$ $`=`$ $`b\left(j1i\stackrel{~}{\omega }{\displaystyle \frac{r_+}{b}}\right)\left(2j72i\stackrel{~}{\omega }{\displaystyle \frac{r_+}{b}}\right)`$ (80) $`+4r_+\stackrel{~}{\omega }(r_+\omega i)+b(4ir_+\omega \mathrm{¯}\lambda ),`$ $`\gamma _j`$ $`=`$ $`\left(j2i\stackrel{~}{\omega }{\displaystyle \frac{r_+}{b}}\right)\left(j5i\stackrel{~}{\omega }{\displaystyle \frac{r_+}{b}}\right)`$ (82) $`+2r_+\omega (2r_+\omega +\stackrel{~}{\omega })+6ib\omega \mathrm{¯}\lambda ,`$ $`\delta _j`$ $`=`$ $`4\omega (r_+\omega +i),`$ (83) $`ϵ_j`$ $`=`$ $`\omega ^2.`$ (84) It is now readily seen that one has noncommuting limits, because $`\alpha _{N+2}`$ $`=`$ $`(N+2)[2imaNa^22(i\omega N/2)]+O(ma^2)`$ (86) $`+O(a^3)+O[(\omega +iN/2)a]+O[(\omega +iN/2)^2]`$ implies that $`c_{N+2}`$ (and hence higher $`c_j`$) are sensitive to the ratio of $`\omega +iN/2`$ and $`a`$. To calculate which are the limiting functions, one has to iterate (77) up to $`c_{N+1}`$, up to the error terms indicated in (LABEL:alphaN2). This in fact is a calculation of the type outlined at the end of Section II A, except that here one has three numerical coefficients in the recursion relation—the lowest-order one plus the leading $`\omega +iN/2`$ (linear) and $`a`$ (linear for $`m0`$, quadratic for $`m=0`$) corrections. In terms of $$h(\mathrm{})=\frac{(\mathrm{}^2m^2)(\mathrm{}^24)^2}{2(\mathrm{}\frac{1}{2})\mathrm{}^3(\mathrm{}+\frac{1}{2})},$$ (88) the required expansion of the separation constant reads $`\mathrm{¯}\lambda `$ $`=`$ $`\mathrm{}(\mathrm{}+1)2(2+{\displaystyle \frac{8}{\mathrm{}(\mathrm{}+1)}})ma\omega `$ (90) $`+[h(\mathrm{}+1)h(\mathrm{})]a^2\omega ^2+O[(a\omega )^3]`$ $`=`$ $`4\frac{10}{3}am\omega +(\frac{10}{21}\frac{10}{189}m^2)a^2\omega ^2`$ (92) $`+O[(a\omega )^3]\text{for }\mathrm{}=2,`$ and for $`\mathrm{}=2`$ some computer algebra yields $$c_{10}=\frac{3633559(4i\omega )+37411744ima+42040064a^2}{1786050(i\omega 4ima+4a^2)},$$ (93) where both the numerator and the denominator are accurate up to the error terms indicated in (LABEL:alphaN2). From (93) it is immediate that $$\underset{\omega \mathrm{\Omega }}{lim}\underset{a0}{lim}c_{10}=\frac{3633559}{1786050}c_{10}^{(3)}.$$ (94) To also find the limit corresponding to (73), substitute (90) into (57), yielding its roots as $`\mathrm{\Omega }(a)`$ $`=`$ $`\mathrm{\Omega }_0+\mathrm{\Omega }_1ma+\mathrm{\Omega }_{20}a^2+\mathrm{\Omega }_{21}m^2a^2`$ (96) $`+O(ma^3)+O(a^4)`$ $`=`$ $`4i+\frac{32}{3}ma+\frac{128}{189}i(65m^218)a^2`$ (98) $`+O(ma^3)+O(a^4)\text{for }\mathrm{}=2,`$ $`\mathrm{\Omega }_0`$ $`=`$ $`\frac{1}{2}iN,`$ (99) $`\mathrm{\Omega }_1`$ $`=`$ $`\frac{2}{3}nN,`$ (100) $`\mathrm{\Omega }_{20}`$ $`=`$ $`\frac{1}{3}inN^2{\displaystyle \frac{(\mathrm{}3)(\mathrm{}+4)}{(2\mathrm{}1)(2\mathrm{}+3)}},`$ (101) $`\mathrm{\Omega }_{21}`$ $`=`$ $`i{\displaystyle \frac{4n^3[21\mathrm{}^2(\mathrm{}+1)^2+2\mathrm{}(\mathrm{}+1)+12]}{27(2\mathrm{}1)(2\mathrm{}+3)}},`$ (102) which in its own right is perhaps a new result for the asymptotics of the special frequencies, agreeing well with the numerical values in for $`\mathrm{}=2`$ and the smallest $`a`$. Note that $`\mathrm{\Omega }_{20}`$ vanishes for $`\mathrm{}=3`$, and has opposite signs for $`\mathrm{}=2`$ and $`\mathrm{}4`$. Using (98) in (93) now yields $$\underset{a0}{lim}\underset{\omega \mathrm{\Omega }(a)}{lim}c_{10}=\frac{69632}{893025}c_{10}^{(1)},$$ (103) where in particular one obtains the same limit both for $`m0`$ and $`m=0`$ even though these two cases are quite different for (93) and (98) individually. Since the normalized function $`limf_2(\omega ,a)`$ is a solution of the Bardeen–Press equation, which is completely specified by its coefficient $`c_{N+2}`$, and since for $`\mathrm{}=2`$ the values (94), (103) are readily checked to correspond to (76) and (73) respectively, the above clarifies the mechanism for the singularity of the Schwarzschild limit, being the zero in $`\alpha _{N+2}`$. ### C Special QNMs We have seen in the preceding subsection that $`f_2(\omega ,a)`$ is wildly varying near $`(\omega =iN/2,a=0)`$. Thus, if near this point $`f_2`$ can tend both to $`A_+\xi _1`$ and $`A_+\xi _3`$ (up to normalization), can it also tend to $`A_+\xi _5`$? After all, the latter is merely a linear superposition of the former two. Thus, one should look for solutions of $$c_{10}(\omega ,a)=c_{10}^{(5)}\left(1\frac{\gamma _5}{\gamma _3}\right)c_{10}^{(1)}+\frac{\gamma _5}{\gamma _3}c_{10}^{(3)}=\frac{2166784}{893025}$$ (104) in the region of validity of (93). Indeed one finds that $`lim_{a0}lim_{\omega \omega _\mathrm{q}(a)}c_{10}(\omega ,a)=c_{10}^{(5)}`$ for $`\omega _\mathrm{q}`$ $`=`$ $`4i{\displaystyle \frac{33078176}{700009}}ma+{\displaystyle \frac{3492608}{41177}}ia^2`$ (106) $`+O(ma^2)+O(a^4).`$ This result has a significant interpretation: to leading order, for $`\omega =\omega _\mathrm{q}(a)`$ the outgoing function into the horizon $`f_2`$ is also outgoing to infinity, so that (106) represents a family of QNMs of (50) for $`(\mathrm{}=2,a0)`$, branching from the special frequency $`\mathrm{\Omega }`$ in the Schwarzschild limit. Clearly, one is interested in the generalization of (106) to arbitrary $`\mathrm{}`$. This can be performed in closed form, since (93), obtained in the above using computer algebra, can be reconstructed from its known limits together with the expansion (LABEL:alphaN2). In fact, considerable cancellation occurs in the answer: while for reference we give the generalizations of (103) and (94) as $$c_{N+2}^{(1)}=\frac{4n^2(8n^5+12n^4+18n^313n^218n45)}{27(2n+3)(N+1)(N+2)!}\left(\frac{N}{2}\right)^N$$ (107) \[from (67) and (73)\] and $$c_{N+2}^{(3)}=c_{N+2}^{(1)}+\frac{6\gamma _3e^N}{(2n+3)^2N(N+1)^2(N+2)}$$ (108) \[from (76) with $`A_+\xi _3=A_+\xi _1+\gamma _3A_+\xi _2`$ and $`A_+\xi _2=6e^{N/2}(r1)^{N/2+2}/[(2n+3)(N+1)(N+2)]+O[(r1)^{N/2+3}]`$ by the explicit form (63) for $`s=2`$ and the three leading coefficients of $`\xi _2`$ as yielded by (13), (14)\] respectively, also without these explicit values one can derive the comparatively simple final result $`\omega _\mathrm{q}`$ $`=`$ $`\frac{1}{2}iN+{\displaystyle \frac{\gamma _5\mathrm{\Omega }_1\gamma _3}{\gamma _5\gamma _3}}ma+{\displaystyle \frac{N\gamma _5+2i\mathrm{\Omega }_{20}\gamma _3}{2(\gamma _5\gamma _3)}}ia^2`$ (110) $`+O(ma^2)+O(a^4).`$ In particular, while the $`O(a)`$-coefficient thus keeps its negative sign for $`\mathrm{}>2`$, the quadratic coefficient, governing the behavior of the special QNMs with $`m=0`$ and positive in (106), is seen to become negative if $`\mathrm{}4`$. For a powerful and instructive check on the preceding, we shall presently rederive the $`m0`$ case (i.e., only the first two terms on the r.h.s.) of (110) from the radial equation (50) for $`s=2`$. Namely, by the Teukolsky–Starobinsky identities this latter equation is physically equivalent to (in particular, has the same QNM spectrum as) the $`s=2`$ equation as long as $`\omega \mathrm{\Omega }(a)`$, a condition which (110) fulfills for $`a0`$. To do so, we investigate in which direction in the $`(\omega ,a)`$-space the QNM (68) can be perturbed so that it continues to satisfy outgoing boundary conditions. In its turn, this calculation buttresses our interpretation of $`𝒫_2^{(0)}`$ (and $`\stackrel{~}{\xi }_1`$ for the ZE) as bona fide QNMs even though they are decreasing on the left; however, this last property will be technically important in the calculation below. Thus, let $`P_{\mathrm{q},2}(a)`$ be the branch of special QNMs, with $`\omega =\omega _\mathrm{q}(a)=iN/2+\omega _{\mathrm{q1}}ma+O(a^2)`$ and $`P_{\mathrm{q},2}(0)=𝒫_2^{(0)}`$, and set $`d_aP_{\mathrm{q},2}(a)|_{a=0}im𝒫_2^{(0)}h_\mathrm{q}`$. Differentiation of (50) yields $`d_rh_\mathrm{q}(r)`$ $`=`$ $`{\displaystyle \frac{r(r1)}{𝒫_{2}^{(0)}{}_{}{}^{2}(r)}}{\displaystyle ^r}𝑑t{\displaystyle \frac{𝒫_{2}^{(0)}{}_{}{}^{2}(t)}{t^3(t1)^3}}B(t),`$ (111) $`B(t)`$ $`=`$ $`N\omega _{\mathrm{q1}}t^4+4\omega _{\mathrm{q1}}t^3+\left({\displaystyle \frac{8n}{3}}6\omega _{\mathrm{q1}}\right)t^2`$ (113) $`+4\left(1n{\displaystyle \frac{n^2}{3}}\right)t2.`$ Our limitation to linear perturbations in $`a`$ enables one to set $`r_+,b=1+O(a^2)`$ etc. in the derivation of (111) and (113), without which these expressions would have been considerably more complicated. For $`P_{\mathrm{q},2}`$ to be outgoing to infinity, on has to set $`^r𝑑t=_{\mathrm{}}^r𝑑t`$ in (111), by a repetition of the arguments in Section V. Since the integrand is entire, one does not need to specify a path; physically this means that our lowest-order calculation is free from branch-cut complications. On the other hand, for outgoing waves into the horizon, one sees from the Born series that $`d_rh(r)`$ should have a residue $`1\omega _{\mathrm{q1}}`$ at $`r=1`$. Writing $`^r𝑑t=K_\mathrm{q}+_1^r𝑑t`$ and expanding the integrand near $`t=1`$, the residue criterion is seen to imply $$K_\mathrm{q}=\frac{8n^2}{9c}(1\omega _{\mathrm{q1}}),$$ (114) with $`c`$ being the residue at $`r=1`$ of $`{\displaystyle \frac{e^{Nr}}{(r1)^{N1}r^3[2(n+1)r3]^2}}={\displaystyle \frac{2ne^{Nr}[(Nn)r+n+1]}{9(r1)^{N+2}r}}`$ (115) $`d_r{\displaystyle \frac{e^{Nr}[2n(4n+1)r^2+3(12n)r3]}{18(r1)^{N+1}r^2[2(n+1)r3]}},`$ (116) that is, $`c=\gamma _3/27`$. For the perturbation to be outgoing to both sides, i.e., to represent a QNM, we now have to solve $`\omega _{\mathrm{q1}}`$ from $$K_\mathrm{q}=_{\mathrm{}}^1𝑑t\frac{𝒫_{2}^{(0)}{}_{}{}^{2}(t)}{t^3(t1)^3}B(t),$$ (117) where a tedious but elementary calculation yields the r.h.s. as $`8n^2(3\omega _{\mathrm{q1}}2nN)/\gamma _5`$. Combination with (114) now gives $`\omega _{\mathrm{q1}}=(\gamma _5\frac{2}{3}nN\gamma _3)/(\gamma _5\gamma _3)`$, i.e., consistent with (110), which is what we set out to show. ## VIII Discussion While over the years the singular nature of the extreme Kerr limit $`a\frac{1}{2}`$ for rotating black holes has been increasingly well understood, the present work shows that also the Schwarzschild limit $`a0`$ is highly delicate. Besides thus pointing out the problems, considerable progress has also been made towards their solution, culminating in (110) for the special QNMs, involving a perhaps beautiful combination of $`\gamma _3`$ (Section III), $`\gamma _5`$ (Section V), and $`\mathrm{\Omega }_1,\mathrm{\Omega }_{20}`$ (Section VII B). However, besides this technical control, it would be desirable to also have a physical understanding. In particular, the total-transmission property of the ‘algebraically special’ solutions is related in to the vanishing of the Weyl scalars $`\mathrm{\Psi }_0`$ or $`\mathrm{\Psi }_4`$ as these occur in the Newman–Penrose formalism of general relativity, an aspect outside this article’s scope. At present it is not yet clear why the argument fails for the Schwarzschild hole. Thus, a Newman–Penrose analysis would be a welcome complement to our wave-equation study on the conceptual side. On the practical side—the calculation of black-hole spectra—the prospects for numerical progress on or near the NIA seem good, even though the present work refutes essentially all numerical claims about the nature of the RWE at $`\mathrm{\Omega }`$ made previously. First, we have proved that the QNM $`\mathrm{\Omega }^{}`$ reported in cannot possibly coincide with $`\mathrm{\Omega }`$. A spurious $`\mathrm{\Omega }^{}`$ could for instance be found if the numerical code ad hoc compensates for an anomalous-point divergence in the $`c_j`$ of (29) which at $`\mathrm{\Omega }`$ actually is absent; however, comparison of the results of Section VII C with the data in in fact suggests the existence of $`\mathrm{\Omega }^{}`$, see below. If an imaginary $`\mathrm{\Omega }^{}`$ would be distinct from, but numerically very close to $`\mathrm{\Omega }`$ (even though it is doubtful whether the data of itself support this conclusion, cf. the Introduction), this would imply that $`\alpha (\omega )`$ has two very nearby zeros, which can be investigated numerically. Second, the analysis of has some serious problems. While the paper does point out (below Table 1) that the pattern of (anti-)Stokes lines near the horizon central to its analysis actually collapses on the imaginary $`\omega `$-axis—where its data are reported—the matter is not pursued to conclude that on that axis analytic continuations in $`r`$ are not suitable to find outgoing waves to start with. Rather, on that axis the generic/anomalous/miraculous distinction is crucial, but the latter cannot be treated by the WKB ideas used in, which are more appropriate for studying $`|\omega |\mathrm{}`$ than for numerics at a finite $`\omega `$. Indeed, Refs. find that (for $`Re\omega 0`$) continuations in $`r`$ can be used as a numerical tool independently of any WKB-type approximations. Remarkably, in each angular-momentum sector $`(\mathrm{},m)`$, of the two one-parameter families—labeled by $`(a,s=\pm 2)`$—of TTMs claimed in, the only one which has been comfirmed numerically (the TTM<sub>R</sub> at $`a=0`$) thus is the one which in fact does not exist! Third, it has been found in Section VII C that in each $`\mathrm{}`$-sector the Schwarzschild special frequency $`\mathrm{\Omega }(0)`$ splits into one $`m`$-multiplet of special QNMs as the rotation is increased from zero. In particular, only for $`m<0`$ do we predict these QNMs to branch into the fourth $`\omega `$-quadrant, in contrast with the claim in. However, it should be noted that the coefficients in (106) are quite large, so that the Schwarzschild limit is approached only for extremely small $`a`$; as yet no numerical data exist for this regime. In fact, it has been shown in Section VII that for the $`s=2`$ radial equation, in this limit the special QNM and TTM<sub>L</sub> become identical (like for the ZE), while for $`s=2`$ the single (in each $`(\mathrm{},m)`$-sector) QNM and TTM<sub>R</sub> cancel each other (like for the RWE). Thus, while correctly cautions for subtleties of the Schwarzschild limit, its precise nature cannot be elucidated from numerics alone. The preceding does not account for the ninth branches of $`(\mathrm{}=2,m=1,2)`$ Kerr QNMs which reports for moderately small $`a`$, and which have slightly smaller $`|Im\omega |`$ than their apparent $`m=1,2`$ counterparts, the latter being accounted for by (106). Since in the Schwarzschild limit all Kerr QNMs have to merge as $`m`$-multiplets by spherical symmetry, this strongly suggests that the former QNMs merge with their $`m<0`$ mirror images into an imaginary $`\mathrm{\Omega }^{}`$, with $`|\mathrm{\Omega }^{}||\mathrm{\Omega }(0)|`$. The $`m=0`$ branches emerging from $`\mathrm{\Omega }(0)`$ and $`\mathrm{\Omega }^{}`$ then supposedly move towards each other along the NIA, until they collide and subsequently move away from this axis. This scenario is consistent with the sign of the $`O(a^2)`$-term in (106); the remark below (110) then hints that $`|\mathrm{\Omega }^{}|>|\mathrm{\Omega }(0)|`$ for $`\mathrm{}4`$. Additional numerical work is needed to explore these fascinating possibilities; also, for a general understanding of the Schwarzschild limit it is unfortunate that the figures in do not show some Kerr modes branching from the high-damping series of Schwarzschild QNMs. A detailed computation of $`\alpha (\omega )`$ as proposed in relation to would also be of more general interest. Namely, in Section VI we have proved that $`\mathrm{\Omega }`$ is the only ‘miraculous’ frequency of the RWE, and the question suggests itself whether it likewise is the only zero of the discontinuity function $`\alpha `$. It turns out that this cannot be the case in general. For small $`\omega `$, one can use the Born analysis of. True to form, the Schwarzschild potentials are an exceptional case (a major conclusion of) in that the leading-order contribution to $`\alpha (\omega )`$ expected generically vanishes. Fortunately the actual leading-order behavior still follows from the first Born approximation, as $$\alpha (\omega )=()^{\mathrm{}+1}2\pi \omega +\text{h.o.t.},$$ (118) incidentally for the ZE as well as for the RWE as required by (26). Eq. (118) can now be compared to the upshot of Section V D, viz., $`{\displaystyle \frac{d[i\alpha (\mathrm{\Omega })]}{d(i\omega )}}`$ $`=`$ $`8\pi n^2\left[2N{\displaystyle \frac{\gamma _3\gamma _5}{\gamma _5^2}}+9N^2\left({\displaystyle \frac{1}{\gamma _4}}{\displaystyle \frac{1}{\gamma _5}}\right)^2\right]`$ (119) $`>`$ $`0.`$ (120) In both calculations one finds that the $`x`$-dependence in $`\alpha (\omega )\delta g(x,\omega )/g(x,\omega )`$ cancels, as it must. Comparison of (118) and (119) now shows that the parity of zeros (counting their multiplicities) of $`\alpha (iz)`$ on $`(0,N/2)`$ is $`()^{\mathrm{}}`$, so that in particular for odd $`\mathrm{}`$ the function $`\alpha (iz)`$ must have at least one zero on the stated interval. Numerical work should urgently verify this last conclusion, as well as (118), (119) in general, also in the context of the claim in that $`Rei\mathrm{\Omega }^{}<N/2`$ for $`\mathrm{}=2`$, but $`>N/2`$ for $`\mathrm{}=3`$. Further, it should be investigated whether WKB methods or those outlined at the end of Section VI can establish the sign of $`i\alpha (\omega )`$ analytically also for $`i\omega \mathrm{}`$. In any case, in view of (26), in the latter limit those signs should be opposite for the RWE and ZE, unlike the situation for $`i\omega 0`$. The way in which throughout this article analytic continuations in $`r`$ and in $`\omega `$ have turned out to be related points to some interesting function theory in $`𝐂^2`$. While these pure-mathematical aspects have not been pursued, as their consequence and almost coincidentally we have arrived at a fairly complete theory of anomalous points and miracles in general, that is, independent of the specifics of the potentials (3), (4). Namely, in practice most potential tails will be either exponential, covered by the Frobenius method in Section II A, or algebraic, covered by the two-step theorem of Section V A. In particular, the latter, while comparatively intractable in other contexts because of their diverging Born series, have turned out not to lead to anomalous points. A similar degree of generality has not yet been achieved for describing merging and canceling modes, as occurring in Section VII, or in for the Pöschl–Teller model. A model-independent study of merging modes is available as the Jordan-block perturbation theory of, but there one starts by assuming the presence of a higher-order mode, which is subsequently split. In particular, it is not known whether the cancellation phenomenon is restricted to long-range potentials. The study of such issues will be facilitated by generalizing to long-range potentials the concepts of a QNM norm and more generally a generalized inner product between two-component vectors (composed of both fields and their associated momenta), which are known to be highly effective for finite-range potentials. The ‘practical’ significance of this work mainly lies in its results on the ‘special’ family of QNMs at small $`a`$, and in the guidance it gives to any subsequent numerical investigations. Namely, there seems to be no reason why the rotation of an astrophysical black hole should ever vanish exactly. Furthermore, even for real-time numerical experiments one does not predict the ‘special’ ZE QNM to be independently observable: since the QNMs are not complete, the highly damped non-oscillating signal of the latter is in principle indistinguishable from the late-time tail caused by the branch cut in $`g`$, consistent with the ZE’s SUSY-equivalence to the RWE, which has been shown in Section V B to have no ‘special’ QNM. However, apart from the dependence of our small-$`a`$ results on the analysis at $`a=0`$, there is no further justification required to study the RWE and ZE closely, given that they have the same status in gravity as the Schrödinger and Dirac equations for the hydrogen atom have in quantum mechanics. Indeed, given its global singularity structure, especially the RWE is the logical next step up in complexity from the hypergeometric equation encountered in the hydrogen problem. In fact, a look at a figure of the Schwarzschild spectrum (e.g.,) shows that the special frequency is completely central, as the place from which the low- and high-damping branches somehow emerge. Hence, one has the hope that an improved understanding of $`\mathrm{\Omega }`$ will lead to further insight into the spectrum in general, for instance by relating the two numerically observed branches to the two potential tails, or more ambitiously, by providing accurate and general asymptotic formulae. ## Acknowledgment I thank P.T. Leung, Y.T. Liu, K.W. Mak, W.M. Suen, H. Verlinde, C.W. Wong and in particular K. Young for discussions and comments on the preprint, and the ITP Beijing as well as the Universities of Amsterdam and Delft for their hospitality. The work is supported in part by the Hong Kong Research Grants Council (grant CUHK 4006/98P). ## A An afterthought on $`\alpha `$ Our aquaintance with $`\alpha `$ has proceeded stepwise. Defined originally as a set of proportionality constants in (9), immediately below at least locally on the NIA the function $`\alpha (\omega )`$ was considered, focusing on the orders of its zeros. Subsequently, in Section VIII the function was studied globally on this axis, and simple ideas of real analysis were used to make a statement about the zeros of $`\alpha `$. The logical elaboration is to continue $`\alpha `$ analytically from the NIA, observing that (9) then yields the continuation of $`g_\mathrm{l}(\omega )`$ ($`g_\mathrm{r}(\omega )`$) to the fourth (third) quadrant and beyond. One attractive property of $`\alpha `$ is that it is invariant under finitely supported perturbations of $`V`$. In fact this last result can be sharpened at once: a perturbation $`e^{\lambda x}`$ cannot change $`\alpha (\omega )`$ for $`|Im\omega |<\lambda /2`$, and subsequently this invariance can be continued to the whole $`\omega `$-plane. Let us presently derive a further important property: if the potential $`V(x)`$ satisfies the decay property of Section V A with $`\beta =\mathrm{}`$ and if in addition it is single-valued near $`x=+\mathrm{}`$ (so that $`V(x>x_0)=_{j=2}^{\mathrm{}}c_jx^j`$), then $`\alpha (\omega )`$ is entire. Thus, for these potentials the definition (9) is even more fortunate than was apparent at first sight: from a two-variable multiple-valued function it creates a one-variable non-branching function. For a proof, supplement $`g(x,\omega )`$ with a branch cut on the negative real $`x`$-axis (in addition to the cut in $`\omega `$ on the NIA), set $`𝒈(x,\omega )(g(x,\omega ),g(x,\omega ))^\mathrm{T}`$, and define the monodromy map by $$𝒈(\mathrm{\Lambda }iϵ,\omega )=(\omega )𝒈(\mathrm{\Lambda }+iϵ,\omega )$$ (A1) ($`ϵ,\mathrm{\Lambda }>0`$). Elementary properties are $`_{jk}(\omega )=_{3j,3k}(\omega )`$ ($`j,k=1,2`$) and $`^{}(\omega )=^1(\omega ^{})`$; this last relation involves a matrix inversion because for complex $`x`$ one has $`g^{}(x,\omega )=g(x^{},\omega ^{})`$, so that under complex conjugation the roles of $`\mathrm{\Lambda }\pm iϵ`$ are reversed. Furthermore, the two-step theorem of Section V A implies that $`_{22}(\omega )=1`$ in the fourth quadrant. To relate $`\alpha `$ to $``$, observe that for a given $`\omega `$, the two-step theorem yields the asymptotic form of $`g`$ for a range of $`\mathrm{arg}x`$ obeying condition (C) of Section V A. Subsequently, $`g`$ can be continued at fixed $`\omega `$ to the physical region $`x>0`$ by use of (A1)—effectively adding a third step to the two-step theorem—and comparison with (9) allows one to read off $`\alpha `$. Thus, for instance for $`Re\omega <0`$ one finds $$\alpha (\omega )=_{12}(\omega ),$$ (A2) making the relation between $`\omega `$\- ($`\alpha `$) and $`x`$-continuations ($``$) maximally explicit. Now let $`\omega =i|\omega |`$, and consider the equality $`g(\mathrm{\Lambda }iϵ,\omega \eta )=_{11}(\omega \eta )g(\mathrm{\Lambda }+iϵ,\omega \eta )+_{12}(\omega \eta )g(\mathrm{\Lambda }+iϵ,\eta \omega )`$ for $`\eta 0`$. By the preceding one has $`_{11}(\omega \eta )=1`$, while in the limit $`g(\mathrm{\Lambda }iϵ,\omega )=g^{}(\mathrm{\Lambda }+iϵ,\omega )`$ and $`g(\mathrm{\Lambda }+iϵ,\eta \omega )g_\mathrm{r}(\mathrm{\Lambda }+iϵ,\omega )`$. Since the latter function is $`e^{i\omega \mathrm{\Lambda }}`$ and thus real for $`ϵ0`$, it follows that $`Relim_{\eta 0}_{12}(\omega \eta )=0`$. Combination with (A2) and the symmetry $`\alpha (\omega ^{})=\alpha ^{}(\omega )`$ then shows that $`lim_{\eta 0}[\alpha (\omega \eta )\alpha (\omega +\eta )]=0`$, so that $`\alpha `$ has no branch point at the origin. Since a Born calculation (cf. (118)) shows that $`\alpha `$ is also bounded near the origin, and since under the stated conditions on $`V`$ the function cannot have singularities for finite $`\omega `$ (cf. Section V A), the derivation is complete. The above is a result of some power: while for algebraic potential tails the large-$`x`$ expansion of $`g`$ furnished by the Born series diverges, one now sees that under the stated conditions on $`V`$ the same series used as a small-$`\omega `$ expansion of $`\alpha `$ necessarily converges. Also, in this situation one has explicit expressions for all branches of $`g`$; for instance, $`g(\omega e^{2\pi i})=[1+\alpha (\omega )\alpha (\omega )]g(\omega )+\alpha (\omega )g(\omega )`$ if $`Re\omega >0`$. Unfortunately, the only available exactly solvable example $`V(x>x_0)=\sigma (\sigma +1)x^2`$ has $`\alpha (\omega )=2i\mathrm{sin}(\pi \sigma )`$, i.e., an $`\omega `$-independent constant which is not particularly instructive. Note that (pending a closer investigation of the convergence of (31)) the result (46) suggests that in the RW case the cut $`\delta g`$ itself is entire, so that $`\alpha `$ would have the same branching properties as $`g`$; there is no contradiction, for (3) is not single-valued at infinity in the $`x`$-variable. For potentials which cause a cut in only one of the outgoing waves (such as (3) and (4)), one has the result that zero-modes can only be located at a zero of $`\alpha `$ on the NIA (Ref. and cf. Section V C). The above development prompts the question whether this result can be generalized to other frequencies. However, the most obvious of such generalizations cannot be true: under finitely supported perturbations of $`V`$, off-axis QNMs will vary continuously while the zeros of $`\alpha `$ will not move at all, so that the two cannot coincide in general. In fact, such QNM perturbations are interesting already on the NIA: since the zero of $`\alpha `$ which coincides with the original QNM will be invariant, it follows that the QNM cannot be shifted along the NIA. Hence, it either has to move away from the axis into a physical branch of the Green’s function, and hence split by symmetry, or move into an unphysical branch and hence effectively disappear. Clearly, in the presence of branch cuts in the outgoing wave(s), the Green’s function on the NIA and beyond into its unphysical branches merits considerable further study, both for its fundamental interest and for the potential of QNM poles on these unphysical branches to cause numerical artifacts.
warning/0001/gr-qc0001089.html
ar5iv
text
# On the gravitational moments of a Dirac particle ## I Introduction Riemann-Cartan geometry arises naturally in the gauge theory of the Poincaré group, see e.g. . One can interpret spacetime coframe $`\vartheta ^\alpha `$ and local Lorentz connection $`\mathrm{\Gamma }_\alpha ^\beta `$ as the gravitational potentials related to the translation group and the Lorentz group, respectively. The two-forms of torsion and curvature, $`T^\alpha :=d\vartheta ^\alpha +\mathrm{\Gamma }_\beta {}_{}{}^{\alpha }\vartheta ^\beta `$, $`R_\alpha {}_{}{}^{\beta }:=d\mathrm{\Gamma }_\alpha {}_{}{}^{\beta }+\mathrm{\Gamma }_\gamma {}_{}{}^{\beta }\mathrm{\Gamma }_\alpha ^\gamma `$, represent the corresponding gauge field strengths. Besides, the Riemann-Cartan spacetime carries a metric $`g_{\alpha \beta }`$ which is covariantly constant: $`Dg_{\alpha \beta }`$ $`=0`$. Usually one chooses $`g_{\alpha \beta }=o_{\alpha \beta }`$, the flat Minkowski metric, thus restricting oneself to orthonormal frames and coframes. General dynamical scheme of the Poincaré gauge theory is well established. The Noether currents of matter fields, $`\mathrm{\Sigma }_\alpha `$ (the canonical energy-momentum three-form) and $`\tau _{\alpha \beta }`$ (the canonical spin three-form), are coupled to the translational $`\vartheta ^\alpha `$ and the Lorentz $`\mathrm{\Gamma }^{\alpha \beta }`$ gauge potentials, respectively. These currents are thus representing the two types of gravitational charges of a matter source. Recently, the Dirac electron theory has been analyzed in flat Minkowski spacetime , and the structure of the canonical energy-momentum and spin currents was studied in detail. It was shown, developing analogy with electrodynamics, that a Dirac particle is naturally characterized by two gravitational moments. In simple physical terms, one can describe them as the Ampére type ring currents induced by the two gravitational charges via spin of a particle. Here we consider the Dirac theory with the gravitational field “switched on”. We demonstrate the consistency of the coupling of the gravitational moments to the torsion and curvature. It is worthwhile to mention that gravitational moments of a Dirac particle were discussed also by Kobzarev and Okun and Khriplovich in the framework of Einstein’s general relativity theory, whereas in gravitational moments of higher spins and in lower dimensions were studied. In this paper, we are using the basic conventions and notations of Bjorken and Drell. In particular, the constant Minkowski metric is $`o_{\alpha \beta }=\mathrm{diag}(+1,1,1,1)`$, and we choose the Dirac matrices $`\gamma ^\alpha `$ in the standard form of . In the exterior algebra on a spacetime, $`,`$ are exterior and interior products, respectively, is the Hodge star operator. Local orthonormal frame $`e_\alpha `$ is dual to a coframe $`\vartheta ^\alpha `$ one-form: $`e_\alpha \vartheta ^\beta =\delta _\alpha ^\beta `$. Finally, starting from the volume one-form $`\eta :={}_{}{}^{}1`$, one defines Trautman’s $`\eta `$-basis as usual by $`\eta _\alpha :=e_\alpha \eta =^{}\vartheta _\alpha `$, $`\eta _{\alpha \beta }:=e_\beta \eta _\alpha =^{}(\vartheta _\alpha \vartheta _\beta )`$ etc. ## II Fermions in the Riemann-Cartan spacetime For the description of classical Dirac particle of mass $`m`$, we will use the formalism of Clifford algebra–valued exterior forms . The following matrix-valued one- or three-forms are basic objects in this approach: $$\gamma :=\gamma _\alpha \vartheta ^\alpha ,{}_{}{}^{}\gamma =\gamma ^\alpha \eta _\alpha .$$ (1) Exterior product yields a two-form: $$\widehat{\sigma }:=\frac{i}{2}\gamma \gamma =\frac{1}{2}\widehat{\sigma }_{\alpha \beta }\vartheta ^\alpha \vartheta ^\beta .$$ (2) The coefficients $`\widehat{\sigma }_{\alpha \beta }:=i\gamma _{[\alpha }\gamma _{\beta ]}`$ generate infinitesimal Lorentz transformations of spinor fields. Defining another constant matrix, $`\gamma _5:=i\gamma ^{\widehat{0}}\gamma ^{\widehat{1}}\gamma ^{\widehat{2}}\gamma ^{\widehat{3}}`$, it is straightforward to prove the fundamental identities for the Clifford algebra-valued objects: $`\widehat{\sigma }_{\alpha \beta }{}_{}{}^{}\widehat{\sigma }`$ $`=`$ $`\eta _{\alpha \beta }i\gamma _5\vartheta _\alpha \vartheta _\beta 2i\vartheta _{[\alpha }e_{\beta ]}^{}\widehat{\sigma },`$ (3) $`{}_{}{}^{}\widehat{\sigma }\widehat{\sigma }_{\alpha \beta }`$ $`=`$ $`\eta _{\alpha \beta }i\gamma _5\vartheta _\alpha \vartheta _\beta +2i\vartheta _{[\alpha }e_{\beta ]}^{}\widehat{\sigma }.`$ (4) The Lagrangian four-form of a Dirac field $`\mathrm{\Psi }`$ is given by $$L_\mathrm{D}=\frac{i}{2}\mathrm{}\left\{\overline{\mathrm{\Psi }}{}_{}{}^{}\gamma D\mathrm{\Psi }+\overline{D\mathrm{\Psi }}{}_{}{}^{}\gamma \mathrm{\Psi }\right\}+{}_{}{}^{}mc\overline{\mathrm{\Psi }}\mathrm{\Psi }.$$ (5) Dirac fields are local sections of the spinor $`SO(1,3)`$-bundle associated with the principal bundle of orthonormal frames. Hence, the spinor covariant derivative in the Riemann-Cartan spacetime is defined by $$D\mathrm{\Psi }:=d\mathrm{\Psi }+\frac{i}{4}\mathrm{\Gamma }^{\alpha \beta }\widehat{\sigma }_{\alpha \beta }\mathrm{\Psi },\overline{D\mathrm{\Psi }}=d\overline{\mathrm{\Psi }}\frac{i}{4}\mathrm{\Gamma }^{\alpha \beta }\overline{\mathrm{\Psi }}\widehat{\sigma }_{\alpha \beta }.$$ (6) The Dirac field equation, which arises from the Lagrangian (5), reads $`i\mathrm{}{}_{}{}^{}\gamma \left(D\mathrm{\Psi }{\scriptscriptstyle \frac{1}{2}}T\mathrm{\Psi }\right)+{}_{}{}^{}mc\mathrm{\Psi }`$ $`=`$ $`0,`$ (7) $`i\mathrm{}\left(\overline{D\mathrm{\Psi }}{\scriptscriptstyle \frac{1}{2}}T\overline{\mathrm{\Psi }}\right){}_{}{}^{}\gamma +{}_{}{}^{}mc\overline{\mathrm{\Psi }}`$ $`=`$ $`0,`$ (8) where $`T:=e_\alpha T^\alpha `$ is the torsion trace one-form. The standard Lagrange–Noether machinery in the gauge gravity, see e.g. , provides a general definition of the energy-momentum and spin currents in non-flat spacetime (accounting also for the possibility of non-minimal coupling and “Pauli-type” terms): $`\mathrm{\Sigma }_\alpha `$ $`:=`$ $`e_\alpha L(e_\alpha D\mathrm{\Psi }^A){\displaystyle \frac{L}{D\mathrm{\Psi }^A}}(e_\alpha \mathrm{\Psi }^A){\displaystyle \frac{L}{\mathrm{\Psi }^A}}`$ (10) $`+D{\displaystyle \frac{L}{T^\alpha }}(e_\alpha T^\beta ){\displaystyle \frac{L}{T^\beta }}(e_\alpha R_\beta {}_{}{}^{\gamma }){\displaystyle \frac{L}{R_\beta ^\gamma }},`$ $`\tau _{\alpha \beta }`$ $`:=`$ $`(\mathrm{}_{\alpha \beta }{}_{B}{}^{A}\mathrm{\Psi }_{}^{B}){\displaystyle \frac{L}{D\mathrm{\Psi }^A}}+\vartheta _{[\alpha }{\displaystyle \frac{L}{T^{\beta ]}}}+D{\displaystyle \frac{L}{R^{\alpha \beta }}}.`$ (11) Here $`\mathrm{\Psi }^A`$ is a set of arbitrary matter fields, with $`\mathrm{}_{\alpha \beta }`$ denoting the corresponding Lorentz group generators. The first and second Noether theorems yield two covariant conservation laws which are fullfilled on the classical matter field equations: $`D\mathrm{\Sigma }_\alpha `$ $`=`$ $`0,`$ (12) $`D\tau _{\alpha \beta }+\vartheta _{[\alpha }\mathrm{\Sigma }_{\beta ]}`$ $`=`$ $`0.`$ (13) In case of the Dirac theory under consideration, matter is described by the pair of independent four-spinors $`\mathrm{\Psi }^A=\{\mathrm{\Psi },\overline{\mathrm{\Psi }}\}`$, and the canonical gravitational currents are straightforwardly obtained after substituting (5) into (10)-(11) \[hereafter $`D_\alpha :=e_\alpha D`$\]: $`\mathrm{\Sigma }_\alpha `$ $`=`$ $`{\displaystyle \frac{i\mathrm{}}{2}}\left(\overline{\mathrm{\Psi }}{}_{}{}^{}\gamma D_\alpha \mathrm{\Psi }D_\alpha \overline{\mathrm{\Psi }}{}_{}{}^{}\gamma \mathrm{\Psi }\right),`$ (14) $`\tau _{\alpha \beta }`$ $`=`$ $`{\displaystyle \frac{\mathrm{}}{4}}\vartheta _\alpha \vartheta _\beta \overline{\mathrm{\Psi }}\gamma \gamma _5\mathrm{\Psi }.`$ (15) ## III Gravitational moments of a Dirac particle In a theory invariant with respect to some internal gauge group $`G`$, the generalized moment is a $`𝒢`$-valued two-form such that its exterior differential produces the polarizational part of the related Noether current, with $`𝒢`$ denoting the Lie algebra of $`G`$. On the Lagrangian level, the moment couples directly to the gauge field strength. To be specific, in the Dirac–Yang-Mills theory the Noether “isospin” current $`J__K=ie\overline{\mathrm{\Psi }}\tau __K\mathrm{\Psi }`$ couples to the gauge potential one-form $`A^_K`$, with $`e`$ denoting the coupling constant, and $`\tau __K`$ being generators of the gauge group. The Gordon decomposition reveals a nontrivial substructure of this coupling by relating the polarization moment two-form $`P__K=i\frac{e\mathrm{}}{2mc}\overline{\mathrm{\Psi }}\tau __K{}_{}{}^{}\widehat{\sigma }\mathrm{\Psi }`$ directly to the gauge field strength $`F^_K`$. One may expect similar results to hold when passing from internal symmetry groups to spacetime symmetries. Geometrically this means a departure from flat Minkowski spacetime by “switching on” gravity. Technically, we can proceed along the same lines as for the Dirac–Yang-Mills theory . We use (7)-(8) in order to express $`\mathrm{\Psi }`$ and $`\overline{\mathrm{\Psi }}`$ in terms of the differentials, and then substitute them back into the Dirac equation and into the Noether currents (14)-(15). After some algebra, we find the squared Dirac equation in the form $$(D{}_{}{}^{}D+2SD+X)\mathrm{\Psi }=0,$$ (16) where the three-form $`S`$ and the four-form $`X`$ read, respectively, $`S`$ $`:=`$ $`{\displaystyle \frac{i}{2}}(D{}_{}{}^{}\widehat{\sigma }+T{}_{}{}^{}\widehat{\sigma }),`$ (17) $`X`$ $`:=`$ $`{}_{}{}^{}\left({\displaystyle \frac{mc}{\mathrm{}}}\right)_{}^{2}+{\displaystyle \frac{1}{4}}{}_{}{}^{}\widehat{\sigma }R_{\alpha \beta }\widehat{\sigma }^{\alpha \beta }`$ (18) $`+{\displaystyle \frac{1}{2}}(d{}_{}{}^{}T`$ $`+`$ $`i{}_{}{}^{}\widehat{\sigma }dT{\displaystyle \frac{1}{2}}T{}_{}{}^{}T+i(D{}_{}{}^{}\widehat{\sigma })T).`$ (19) Here we have used the Ricci identity for the spinor covariant derivative (6): $$DD\mathrm{\Psi }=\frac{i}{4}R_{\alpha \beta }\widehat{\sigma }^{\alpha \beta }\mathrm{\Psi }.$$ (20) One can immediately verify that equation (16) can be derived from the Lagrange four-form $$L_{D^2}=L^{(c)}+L^{(p)},$$ (21) where the convective Lagrangian and the polarizational Lagrangian read, respectively: $`L^{(c)}`$ $`:=`$ $`{\displaystyle \frac{1}{2}}\left({\displaystyle \frac{\mathrm{}^2}{mc}}{}_{}{}^{}\overline{D\mathrm{\Psi }}D\mathrm{\Psi }+{}_{}{}^{}mc\overline{\mathrm{\Psi }}\mathrm{\Psi }\right),`$ (22) $`L^{(p)}`$ $`=`$ $`M_{\alpha \beta }R^{\alpha \beta }+M_\alpha T^\alpha {\displaystyle \frac{\mathrm{}^2}{8mc}}T{}_{}{}^{}T\overline{\mathrm{\Psi }}\mathrm{\Psi }.`$ (23) Here $`M_{\alpha \beta }`$ is the Lorentz gravitational moment given by $$M_{\alpha \beta }:=\frac{\mathrm{}^2}{16mc}\overline{\mathrm{\Psi }}({}_{}{}^{}\widehat{\sigma }\widehat{\sigma }_{\alpha \beta }+\widehat{\sigma }_{\alpha \beta }{}_{}{}^{}\widehat{\sigma })\mathrm{\Psi },$$ (24) and $$M_\alpha :=\frac{\mathrm{}^2}{4mc}[i\overline{\mathrm{\Psi }}{}_{}{}^{}\widehat{\sigma }D_\alpha \mathrm{\Psi }iD_\alpha \overline{\mathrm{\Psi }}{}_{}{}^{}\widehat{\sigma }\mathrm{\Psi }(e_\alpha {}_{}{}^{}\overline{D\mathrm{\Psi }})\mathrm{\Psi }\overline{\mathrm{\Psi }}(e_\alpha {}_{}{}^{}D\mathrm{\Psi })].$$ (25) We are now in a position to find the decosmposition of the energy-momentum and spin currents into the convective and polarization parts corresponding to the decomposition of the Lagrangian (21), $`\mathrm{\Sigma }_\alpha `$ $`=`$ $`\mathrm{\Sigma }_\alpha ^{(c)}+\mathrm{\Sigma }_\alpha ^{(p)},`$ (26) $`\tau _{\alpha \beta }`$ $`=`$ $`\tau _{\alpha \beta }^{(c)}+\tau _{\alpha \beta }^{(p)}.`$ (27) A straightforward use of the convective Lagrangian (22) in (10) and (11) yields the the three-forms $`\mathrm{\Sigma }_\alpha ^{(c)}`$ $`:=`$ $`{\displaystyle \frac{mc}{2}}\overline{\mathrm{\Psi }}\mathrm{\Psi }\eta _\alpha +{\displaystyle \frac{\mathrm{}^2}{4mc}}[{}_{}{}^{}(\overline{D\mathrm{\Psi }})D_\alpha \mathrm{\Psi }+D_\alpha \overline{\mathrm{\Psi }}{}_{}{}^{}D\mathrm{\Psi }`$ (29) $`+(e_\alpha {}_{}{}^{}\overline{D\mathrm{\Psi }})D\mathrm{\Psi }+\overline{D\mathrm{\Psi }}(e_\alpha {}_{}{}^{}D\mathrm{\Psi })],`$ $`\tau _{\alpha \beta }^{(c)}`$ $`=`$ $`{\displaystyle \frac{i\mathrm{}^2}{8mc}}\left({}_{}{}^{}\overline{D\mathrm{\Psi }}\widehat{\sigma }_{\alpha \beta }\mathrm{\Psi }\overline{\mathrm{\Psi }}\widehat{\sigma }_{\alpha \beta }{}_{}{}^{}D\mathrm{\Psi }\right).`$ (30) For the polarizational part, we need the derivatives with respect to curvature and torsion: $`{\displaystyle \frac{L^{(p)}}{R^{\alpha \beta }}}`$ $`=`$ $`M_{\alpha \beta },`$ (31) $`{\displaystyle \frac{L^{(p)}}{T^\alpha }}`$ $`=`$ $`M_\alpha +{\displaystyle \frac{\mathrm{}^2}{4mc}}(e_\alpha {}_{}{}^{}T)\overline{\mathrm{\Psi }}\mathrm{\Psi }=:\text{}_\alpha .`$ (32) Here we recover the correct translational grativational moment $$\text{}_\alpha =\frac{i\mathrm{}^2}{4mc}[\overline{\mathrm{\Psi }}(e_\alpha {}_{}{}^{}\widehat{\sigma })D\mathrm{\Psi }+\overline{D\mathrm{\Psi }}(e_\alpha {}_{}{}^{}\widehat{\sigma })\mathrm{\Psi }].$$ (33) In (32) we used the Riemann-Cartan the identity which holds for all spinor fields satisfying the Dirac equation: $$i\left(\overline{\mathrm{\Psi }}^{}\widehat{\sigma }D\mathrm{\Psi }\overline{D\mathrm{\Psi }}^{}\widehat{\sigma }\mathrm{\Psi }\right)={}_{}{}^{}D(\overline{\mathrm{\Psi }}\mathrm{\Psi }){}_{}{}^{}T\overline{\mathrm{\Psi }}\mathrm{\Psi }.$$ (34) Substituting (31),(32) into (10),(11), we finally obtain the polarizational energy-momentum and spin currents of a Dirac particle in the Riemann-Cartan spacetime: $`\mathrm{\Sigma }_\alpha ^{(p)}`$ $`=`$ $`D\text{}_\alpha +\mathrm{\Sigma }_\alpha ^{(RC)},`$ (35) $`\tau _{\alpha \beta }^{(p)}`$ $`=`$ $`\vartheta _{[\alpha }\text{}_{\beta ]}+DM_{\alpha \beta }+\tau _{\alpha \beta }^{(RC)}.`$ (36) The terms $`\mathrm{\Sigma }_\alpha ^{(RC)}`$ and $`\tau _{\alpha \beta }^{(RC)}`$ contain curvature and torsion explicitly; these contributions are absent in flat Minkowski spacetime. For completeness, we give their explicit form: $`\mathrm{\Sigma }_\alpha ^{(RC)}`$ $`=`$ $`(e_\alpha M_{\rho \sigma })R^{\rho \sigma }\text{}_\alpha T+(e_\alpha \text{}_\beta e_\beta \text{}_\alpha )T^\beta `$ (39) $`(e_\alpha e_\beta U)T^\beta +(e_\alpha U)T+(e_\alpha {}_{}{}^{}U)^{}T`$ $`{\displaystyle \frac{\mathrm{}^2}{8mc}}[(e_\alpha T){}_{}{}^{}T+Te_\alpha {}_{}{}^{}T]\overline{\mathrm{\Psi }}\mathrm{\Psi },`$ $`\tau _{\alpha \beta }^{(RC)}`$ $`=`$ $`e_\gamma (M_{\alpha \beta }T^\gamma ).`$ (40) As was noticed in , the three-form $$U=\frac{1}{2}\vartheta ^\alpha \stackrel{ˇ}{M}_\alpha =\frac{i\mathrm{}^2}{4mc}\left(\overline{\mathrm{\Psi }}^{}\widehat{\sigma }D\mathrm{\Psi }\overline{D\mathrm{\Psi }}^{}\widehat{\sigma }\mathrm{\Psi }\right)$$ (41) plays a role of a “superpotential” from which all the moments of a Dirac particle are generated. For comparison, it is instructive to collect the properties of the moments in a Table I. ## IV Discussion and conclusion In this paper, we have generalized our previous discussion of the gravitational moments to the case of the Riemann-Cartan spacetime. Although the structure of the torsion- and curvature-dependent terms (39) and (40) is not physically transparent, it is remarkable that the gravitational moments $`\text{}_\alpha `$ and $`M_{\alpha \beta }`$ are the same for a Minkowski and a Riemann-Cartan spacetime. Moreover, the specific coupling $`\{moment\times gaugefieldstrength\}`$ on the Lagrangian level is correctly reproduced in the Poincaré gauge–Dirac theory, see (23). The results obtained are helpful for deepening our understanding of the classical limit of the Dirac theory. A relevant general discussion of the low-energy limit can be found in (for non-inertial frames in flat spacetime), and in (for the Riemann-Cartan spacetime). The present study clearly demonstrates the complete consistency of the properties of two gravitational moments with the fundamental structure of the Poincaré gauge gravity which naturally operates with the two types of gravitational charge, mass and spin. At the same time, the definition and properties of a gravitational moment in a purely Riemannian spacetime of Einstein’s general relativity (GR) remain unclear. A somewhat paradoxical situation arises: only a translational gravitational charge (mass, or energy-momentum) is available in GR, but evidently only a Lorentz gravitational moment can survive for the case of the vanishing torsion \[recall that a moment is “conjugated” to a corresponding gauge field strength, cf. (31)-(32)\]. This problem will be analyzed elsewhere. ## V Acknowledgements This work was supported by Deutsche Forschungsgemeinschaft (Bonn) under project He 528/17-2.
warning/0001/hep-ph0001057.html
ar5iv
text
# References The absolutely dominant processes in weak decays of the $`b`$ and $`c`$ hadrons are, naturally, those associated with the decay of the heavy quark. It is with these decays where, understandably, lies the main interest of the current phenomenological studies, with the prospects for precision determination of the CKM mixing matrix and for uncovering a CP violation in $`B`$ mesons. The present paper however deals with a rather sub-dominant type of decay of strange heavy hyperons into non-strange ones, a process closely analogous to the decays of ordinary strange hyperons, and associated with the decay of the strange, rather than heavy, quark. These decays are interesting for at least two reasons: one is that the branching fractions of the decays $`\mathrm{\Xi }_b\mathrm{\Lambda }_b\pi `$ and $`\mathrm{\Xi }_c\mathrm{\Lambda }_c\pi `$ are not hopelessly small and one may expect that these decays can be studied experimentally, inasmuch as it will be feasible to study any exclusive nonleptonic decays of the $`b`$ and $`c`$ cascade hyperons<sup>1</sup><sup>1</sup>1The branching ratio $`B(\mathrm{\Xi }_b^{}\mathrm{\Lambda }_b\pi ^{})`$ may well exceed 1%, thus possibly being the largest among the branching fractions for individual exclusive nonleptonic decay channels of the $`\mathrm{\Xi }_b^{}`$., and the other reason being that these decays provide a case for a study of the ‘old’ physics in a new setting, namely a study of the structure of baryons containing one heavy quark. Thus these decays offer a testing ground for a combination of the ‘older’ methods, such as the flavor SU(3) symmetry and the current algebra with the ‘newer’ theoretical ideas related to the heavy quark limit. Moreover, as will be shown, the difference between the amplitudes of the decays $`\mathrm{\Xi }_c\mathrm{\Lambda }_c\pi `$ and $`\mathrm{\Xi }_b\mathrm{\Lambda }_b\pi `$ is related, through PCAC and the SU(3) symmetry, to the matrix elements of four-quark operators, that govern, within the heavy quark expansion, the differences of the total inclusive weak decay rates within the triplets of the heavy baryons: ($`\mathrm{\Xi }_c,\mathrm{\Lambda }_c`$) and ($`\mathrm{\Xi }_b,\mathrm{\Lambda }_b`$). The latter matrix elements are a crucial ingredient in understanding the pre-asymptotic effects in the inclusive decay rates of heavy baryons, which are discussed with a recently renewed interest in relation to the data on $`\tau (\mathrm{\Lambda }_b)/\tau (B^0)`$ (for a most recent mini-review see Ref., see also the recent papers ). A very approximate estimate of the absolute rates of the discussed decays can be done by comparing them to similar strangeness-changing decays of ordinary strange hyperons, with rates typically of order $`0.01ps^1`$. For the charmed hyperons the mass difference between $`\mathrm{\Xi }_c`$ and $`\mathrm{\Lambda }_c`$ is known to be about 180 MeV, i.e. quite close to the mass differences of ordinary hyperons differing by one unit of strangeness, and the mass difference between $`\mathrm{\Xi }_b`$ and $`\mathrm{\Lambda }_b`$ should be very close to that for the charmed hyperons due to the heavy quark limit considerations: $$M(\mathrm{\Xi }_b)M(\mathrm{\Lambda }_b)=M(\mathrm{\Xi }_c)M(\mathrm{\Lambda }_c)+O(m_c^2m_b^2),$$ (1) since there are no terms of order $`1/m_Q`$ in this mass splitting. In a more detailed consideration, to be discussed in this paper, the baryonic matrix elements, determining the decay amplitudes, are somewhat different from those for decays of ordinary hyperon decays, thus the specific absolute rates can differ from the simplistic estimate. However, using the mentioned relation of the difference of the amplitudes for $`\mathrm{\Xi }_c`$ and $`\mathrm{\Xi }_b`$ decays to the difference of the lifetimes of the two $`\mathrm{\Xi }_c`$ hyperons and the $`\mathrm{\Lambda }_c`$, we find that this difference alone would correspond to the rate of, e.g. the decay $`\mathrm{\Xi }_c^0\mathrm{\Lambda }_c\pi ^{}`$, equal to $`9\times 10^3ps^1`$, i.e. in the same range as the simplistic estimate. Thus comparing these estimates for the absolute rates with the known lifetimes of the $`\mathrm{\Xi }_c`$ and $`\mathrm{\Xi }_b`$ hyperons, one concludes that the discussed decays should have branching fractions at a per mill level for the $`\mathrm{\Xi }_c`$ hyperons and at a one percent level for the $`\mathrm{\Xi }_b`$. In addition to the strangeness changing decays of heavy cascade hyperons with emission of a pion, we consider here, as a ‘by-product’, two other types of $`\mathrm{\Xi }_Q\mathrm{\Lambda }_Q`$ transitions: with emission of a lepton pair and with emission of a photon, which both of course have much lower probability than the pion transitions. For the semileptonic transitions we note that the axial form factor of the weak current is zero in the heavy quark limit for both $`\mathrm{\Xi }_b\mathrm{\Lambda }_b\mathrm{}\nu `$ and $`\mathrm{\Xi }_c\mathrm{\Lambda }_c\mathrm{}\nu `$, while the vector form factor is $`g_V=1`$ in the flavor SU(3) limit (with the usual further consequences of the Ademollo-Gatto theorem for the SU(3) violation effects). For the photon transitions it is concluded here that the decay $`\mathrm{\Xi }_b\mathrm{\Lambda }_b\gamma `$ is forbidden in the heavy quark limit, while this generally is not the case for $`\mathrm{\Xi }_c\mathrm{\Lambda }_c\gamma `$. The discussed transitions among the heavy hyperons are induced by two underlying weak processes: the ‘spectator’ decay of a strange quark, $`su\overline{u}d`$, $`su\mathrm{}\nu `$, or $`sd\gamma `$, which does not involve the heavy quark, and the ‘non-spectator’ weak scattering (WS) $$sccd$$ (2) trough the weak interaction of the $`cd`$ and $`sc`$ currents. One can also readily see that the WS mechanism contributes only to the decays $`\mathrm{\Xi }_c\mathrm{\Lambda }_c\pi `$ and generally, through a photon emission in WS, to the radiative transition $`\mathrm{\Xi }_c^+\mathrm{\Lambda }_c\gamma `$, while the semileptonic decay $`\mathrm{\Xi }_c^0\mathrm{\Lambda }_c\mathrm{}\nu `$ and all the transitions between the $`\mathrm{\Xi }_b`$ hyperons and the $`\mathrm{\Lambda }_b`$ are contributed only by the ‘spectator’ processes. An important starting point in considering the transitions $`\mathrm{\Xi }_Q\mathrm{\Lambda }_Q`$ induced by the ‘spectator’ decay of the strange quark, is that in the heavy quark limit the spin of the heavy quark completely decouples from the spin variables of the light component of the baryon, and that the latter light component in both the initial and the final baryon forms a $`J^P=0^+`$ state with the quantum numbers of a diquark. Thus these transitions are of a $`0^+0^+`$ type, which imposes strong constraints on the decay amplitudes. In particular, for the pion emission, these constraints imply that the decay amplitude is purely $`S`$ wave, while the $`P`$ wave amplitude is zero in the limit of infinite heavy quark mass<sup>2</sup><sup>2</sup>2For a general phenomenological treatment of the amplitudes of hyperon pion transitions, $`B^{}B\pi `$, in terms of partial waves, see e.g. the textbook .. The implication for the ‘spectator’ radiative transition is that it is forbidden in this limit, thus predicting a strong suppression of the decay $`\mathrm{\Xi }_b^0\mathrm{\Lambda }_b\gamma `$. For the semileptonic decays, contributed only by the ‘spectator’ decay, the constraint from the $`0^+0^+`$ transition is that the axial hadronic form factor is vanishing, $`g_A=0`$, while the vector form factor is $`g_V=1`$. The corresponding decay rate $`\mathrm{\Gamma }(\mathrm{\Xi }_Q\mathrm{\Lambda }_Qe\nu )=G_F^2\mathrm{sin}^2\theta _c(\mathrm{\Delta }M)^5/(60\pi ^3)1.0\times 10^6s^1`$ is however too small to be of a possible phenomenological relevance in the nearest future. For further consideration of the pion transitions $`\mathrm{\Xi }_Q\mathrm{\Lambda }_Q\pi `$ we write the well known expression for the nonleptonic strangeness-changing weak Hamiltonian (see e.g. ) $`H_W=`$ $`\sqrt{2}G_F\mathrm{sin}\theta _c\{(C_++C_{})[(\overline{u}_L\gamma _\mu s_L)(\overline{d}_L\gamma _\mu u_L)(\overline{c}_L\gamma _\mu s_L)(\overline{d}_L\gamma _\mu c_L)]+`$ (3) $`(C_+C_{})[(\overline{d}_L\gamma _\mu s_L)(\overline{u}_L\gamma _\mu u_L)(\overline{d}_L\gamma _\mu s_L)(\overline{c}_L\gamma _\mu c_L)]\}.`$ In this formula the weak Hamiltonian is assumed to be normalized (in LLO) at $`\mu =m_c`$, so that the renormalization coefficients are $`C_{}=C_+^2=\left(\alpha _s(m_c)/\alpha _s(m_W)\right)^{12/25}`$. The terms in the Hamiltonian (3) without the charmed quark fields describe the ‘spectator’ nonleptonic decay of the strange quark, while those with the $`c`$ quark correspond to the WS process (2). It should be noted that the part of $`H_W`$ with (virtual) charmed quarks indirectly contributes to the ‘spectator’ process as well by providing a GIM cutoff for the ‘penguin’ mechanism at $`\mu =m_c`$. However at any normalization point $`\mu `$ below $`m_c`$ this part does not explicitly show up, and for this reason we refer to the terms of $`H_W`$ without the charmed quark fields as ‘spectator’ ones and those with $`c`$ and $`\overline{c}`$ as ‘non-spectator’ ones. One could evolve the weak Hamiltonian down to a low normalization point $`\mu `$, such that $`\mu m_c`$ to make this separation explicit (the ‘spectator’ part then evolves according to the treatment in Ref., while the evolution of the ‘non-spectator’ part is described by the ‘hybrid’ anomalous dimension , and is essentially equivalent to that considered in ). However, the present paper makes no attempt at constructing models for the hadronic matrix elements at a low $`\mu `$, thus writing the corresponding formulas here would be redundant and it is quite sufficient for our purposes to use the expression (3) at $`\mu =m_c`$ with the separation between the ‘spectator’ and the ‘non-spectator’ parts as noted. As discussed above, the ‘spectator’ process gives rise only to the $`S`$ wave amplitudes of the decays $`\mathrm{\Xi }_Q\mathrm{\Lambda }_Q\pi `$, while the ‘non-spectator’ part involves the spin of the charmed quark, and generally may induce a $`P`$ wave as well as an $`S`$ wave in the decays of the $`\mathrm{\Xi }_c`$ hyperons. According to the well known current algebra technique, the $`S`$ wave amplitudes of pion emission can be considered in the chiral limit at zero four-momentum of the pion, where they are described by the PCAC reduction formula (pole terms are absent in these processes): $$\mathrm{\Lambda }_Q\pi _i(p=0)|H_W|\mathrm{\Xi }_Q=\frac{\sqrt{2}}{f_\pi }\mathrm{\Lambda }_Q|[Q_i^5,H_W]|\mathrm{\Xi }_Q,$$ (4) where $`\pi _i`$ is the pion triplet in the Cartesian notation, and $`Q_i^5`$ is the corresponding isotopic triplet of axial charges. The constant $`f_\pi 130MeV`$, normalized by the charged pion decay, is used here, hence the coefficient $`\sqrt{2}`$ in eq.(4). It is straightforward to see from eq.(4) that in the PCAC limit the discussed decays should obey the $`\mathrm{\Delta }I=1/2`$ rule. Indeed, the commutator of the weak Hamiltonian with the axial charges transforms under the isotopic SU(2) in the same way as the Hamiltonian itself. In other words, the $`\mathrm{\Delta }I=1/2`$ part of $`H_W`$ after the commutation gives an $`\mathrm{\Delta }I=1/2`$ operator, while the $`\mathrm{\Delta }I=3/2`$ part after the commutation gives an $`\mathrm{\Delta }I=3/2`$ operator. The latter operator however cannot have a non vanishing matrix element between an isotopic singlet, $`\mathrm{\Lambda }_Q`$, and an isotopic doublet, $`\mathrm{\Xi }_Q`$. Thus the $`\mathrm{\Delta }I=3/2`$ part of $`H_W`$ gives no contribution to the $`S`$ wave amplitudes in the PCAC limit. The $`P`$ wave part of the amplitude vanishes at a zero pion momentum and thus is not given by eq.(4). However, as discussed, the $`P`$ wave can arise only in the decays of charmed hyperons and only from the ‘non-spectator’ part of the $`H_W`$, which is pure $`\mathrm{\Delta }I=1/2`$ to start with. Once the isotopic properties of the decay amplitudes are fixed, one can concentrate on specific charge decay channels, e.g. $`\mathrm{\Xi }_b^{}\mathrm{\Lambda }_b\pi ^{}`$ and $`\mathrm{\Xi }_c^0\mathrm{\Lambda }_c\pi ^{}`$. An application of the PCAC relation (4) with the Hamiltonian from eq.(3) to these decays, gives the expressions for the amplitudes at $`p=0`$ in terms of baryonic matrix elements of four-quark operators: $`\mathrm{\Lambda }_b\pi ^{}(p=0)|H_W|\mathrm{\Xi }_b^{}=`$ $`{\displaystyle \frac{\sqrt{2}}{f_\pi }}G_F\mathrm{sin}\theta _c\mathrm{\Lambda }_b|\left(C_++C_{}\right)\left[(\overline{u}_L\gamma _\mu s_L)(\overline{d}_L\gamma _\mu d_L)(\overline{u}_L\gamma _\mu s_L)(\overline{u}_L\gamma _\mu u_L)\right]+`$ $`\left(C_+C_{}\right)\left[(\overline{d}_L\gamma _\mu s_L)(\overline{u}_L\gamma _\mu d_L)(\overline{u}_L\gamma _\mu s_L)(\overline{u}_L\gamma _\mu u_L)\right]|\mathrm{\Xi }_b^{}=`$ $`{\displaystyle \frac{\sqrt{2}}{f_\pi }}G_F\mathrm{sin}\theta _c\mathrm{\Lambda }_b|C_{}\left[(\overline{u}_L\gamma _\mu s_L)(\overline{d}_L\gamma _\mu d_L)(\overline{d}_L\gamma _\mu s_L)(\overline{u}_L\gamma _\mu d_L)\right]`$ $`{\displaystyle \frac{C_+}{3}}\left[(\overline{u}_L\gamma _\mu s_L)(\overline{d}_L\gamma _\mu d_L)+(\overline{d}_L\gamma _\mu s_L)(\overline{u}_L\gamma _\mu d_L)+2(\overline{u}_L\gamma _\mu s_L)(\overline{u}_L\gamma _\mu u_L)\right]|\mathrm{\Xi }_b^{},`$ (5) where in the last transition the operator structure with $`\mathrm{\Delta }I=3/2`$ giving a vanishing contribution is removed and only the structures with explicitly $`\mathrm{\Delta }I=1/2`$ are retained, and $`\mathrm{\Lambda }_c\pi ^{}(p=0)|H_W|\mathrm{\Xi }_c^0=\mathrm{\Lambda }_b\pi ^{}(p=0)|H_W|\mathrm{\Xi }_b^{}+`$ $`{\displaystyle \frac{\sqrt{2}}{f_\pi }}G_F\mathrm{sin}\theta _c\mathrm{\Lambda }_c|\left(C_++C_{}\right)(\overline{c}_L\gamma _\mu s_L)(\overline{u}_L\gamma _\mu c_L)+`$ $`\left(C_+C_{}\right)(\overline{u}_L\gamma _\mu s_L)(\overline{c}_L\gamma _\mu c_L)|\mathrm{\Xi }_c^0.`$ (6) In the latter formula the first term on the r.h.s. expresses the fact that in the heavy quark limit the ‘spectator’ amplitudes do not depend on the flavor or the mass of the heavy quark<sup>3</sup><sup>3</sup>3The non-relativistic normalization for the heavy quark field is used here, corresponding to $`Q|Q^{}Q|Q=1`$. Thus the amplitudes do not contain normalization factors related to the heavy quark mass.. The rest of the expression (6) describes the ‘non-spectator’ contribution to the $`S`$ wave of the charmed hyperon decay. Using the flavor SU(3) symmetry this contribution can be related to the difference of lifetimes within the charmed hyperon triplet as follows. Due to the absence of correlation of the spin of the heavy quark in the hyperons with its light ‘environment’, the terms, involving the axial current of the charmed quark in the operators in eq.(6), give no contribution to the matrix elements. Thus the only relevant matrix elements are $$\mathrm{\Lambda }_c|(\overline{c}\gamma _\mu c)(\overline{u}\gamma _\mu s)|\mathrm{\Xi }_c^0=x\mathrm{and}\mathrm{\Lambda }_c|(\overline{c}_i\gamma _\mu c_k)(\overline{u}_k\gamma _\mu s_i)|\mathrm{\Xi }_c^0=y,$$ (7) where, by the SU(3) symmetry, the quantities $`x`$ and $`y`$ coincide with those introduced in in terms of differences of diagonal matrix elements over the hyperons: $`x={\displaystyle \frac{1}{2}}(\overline{c}\gamma _\mu c)\left[(\overline{u}\gamma _\mu u)(\overline{s}\gamma _\mu s)\right]_{\mathrm{\Xi }_c^0\mathrm{\Lambda }_c}={\displaystyle \frac{1}{2}}(\overline{c}\gamma _\mu c)\left[(\overline{s}\gamma _\mu s)(\overline{d}\gamma _\mu d)\right]_{\mathrm{\Lambda }_c\mathrm{\Xi }_c^+},`$ (8) $`y={\displaystyle \frac{1}{2}}(\overline{c}_i\gamma _\mu c_k)\left[(\overline{u}_k\gamma _\mu u_i)(\overline{s}_k\gamma _\mu s_i)\right]_{\mathrm{\Xi }_c^0\mathrm{\Lambda }_c}={\displaystyle \frac{1}{2}}(\overline{c}_i\gamma _\mu c_k)\left[(\overline{s}_k\gamma _\mu s_i)(\overline{d}_k\gamma _\mu d_i)\right]_{\mathrm{\Lambda }_c\mathrm{\Xi }_c^+}`$ with the notation for the differences of the matrix elements: $`𝒪_{AB}=A|𝒪|AB|𝒪|B`$. In terms of the quantities $`x`$ and $`y`$ in eq.(7) the difference of the $`S`$ wave decay amplitudes from eq.(6) is written as $`\mathrm{\Delta }A_S\mathrm{\Lambda }_c\pi ^{}(p=0)|H_W|\mathrm{\Xi }_c^0\mathrm{\Lambda }_b\pi ^{}(p=0)|H_W|\mathrm{\Xi }_b^{}=`$ $`{\displaystyle \frac{G_F\mathrm{sin}\theta _c}{2\sqrt{2}f_\pi }}\left[\left(C_{}C_+\right)x\left(C_++C_{}\right)y\right].`$ (9) On the other hand, the same quantities defined by eqs.(8) describe within the heavy quark expansion the differences of the inclusive weak decay rates within the triplet of the charmed baryons . These quantities were extracted from the current data on the lifetime differences for the charmed baryons. In particular it was found that the naive quark model relation $`x=y`$ between the ($`\mu `$ independent) $`x`$ and the ($`\mu `$ dependent $`y`$) does not hold at any $`\mu `$ below $`m_c`$. The numerical value of $`x`$ is found as $`x=(0.04\pm 0.01)GeV^3`$, while the value of $`y`$ at $`\mu =m_c`$ is found to be $`y=0.019\pm 0.009GeV^3`$, which values can be directly used in eq.(9). Because of a correlation in the errors in these estimates and for possible future reference to (hopefully) more precise future data on the lifetimes, it is rather appropriate to express the difference of the amplitudes $`\mathrm{\Delta }A_S`$ in eq.(9) directly in terms of the lifetimes of the charmed hyperons, using the formulas from Ref.. In terms of the operators normalized at $`\mu =m_c`$ the relations for the differences of the total decay rates, including the dominant Cabibbo-unsuppressed nonleptonic decays as well as the decays with single Cabibbo suppression and the semileptonic decays, read as $`\mathrm{\Gamma }(\mathrm{\Xi }_c^0)\mathrm{\Gamma }(\mathrm{\Lambda }_c)={\displaystyle \frac{G_F^2m_c^2}{4\pi }}\mathrm{cos}^2\theta _c\{x[\mathrm{cos}^2\theta _cC_+C_{}+{\displaystyle \frac{\mathrm{sin}^2\theta _c}{4}}(6C_+C_{}+5C_+^2+5C_{}^2)]+`$ $`y[3\mathrm{cos}^2\theta _cC_+C_{}+{\displaystyle \frac{3\mathrm{sin}^2\theta _c}{4}}(6C_+C_{}3C_+^2+C_{}^2)+2]\}{\displaystyle \frac{G_F^2m_c^2}{4\pi }}(1.39x+5.56y),`$ $`\mathrm{\Gamma }(\mathrm{\Lambda }_c)\mathrm{\Gamma }(\mathrm{\Xi }_c^+)={\displaystyle \frac{G_F^2m_c^2}{4\pi }}\{x{\displaystyle \frac{\mathrm{cos}^4\theta _c}{4}}(5C_+^2+5C_{}^22C_+C_{})+`$ (10) $`y[{\displaystyle \frac{3\mathrm{cos}^4\theta _c}{4}}(C_{}^23C_+^22C_+C_{})2(\mathrm{cos}^2\theta _c\mathrm{sin}^2\theta _c)]\}{\displaystyle \frac{G_F^2m_c^2}{4\pi }}(2.88x+3.16y),`$ where for simplification of the subsequent relation the numerical values are substituted for the coefficients $`C_+`$ and $`C_{}`$: $`C_+=0.80`$, $`C_{}=1.55`$, corresponding to $`\alpha _s(m_c)/\alpha _s(m_W)=2.5`$. The relations (10) allow to eliminate the quantities $`x`$ and $`y`$ from the expression (9) in favor of the differences of the total decay rates: $`\mathrm{\Delta }A_S{\displaystyle \frac{\sqrt{2}\pi \mathrm{sin}\theta _c}{G_Fm_c^2f_\pi }}\left[0.45\left(\mathrm{\Gamma }(\mathrm{\Xi }_c^0)\mathrm{\Gamma }(\mathrm{\Lambda }_c)\right)+0.04\left(\mathrm{\Gamma }(\mathrm{\Lambda }_c)\mathrm{\Gamma }(\mathrm{\Xi }_c^+)\right)\right]=`$ $`10^7\left[0.97\left(\mathrm{\Gamma }(\mathrm{\Xi }_c^0)\mathrm{\Gamma }(\mathrm{\Lambda }_c)\right)+0.09\left(\mathrm{\Gamma }(\mathrm{\Lambda }_c)\mathrm{\Gamma }(\mathrm{\Xi }_c^+)\right)\right]\left({\displaystyle \frac{1.4GeV}{m_c}}\right)^2ps,`$ (11) where, clearly, in the latter form the widths are assumed to be expressed in $`ps^1`$, and $`m_c=1.4GeV`$ is used as a ‘reference’ value for the charmed quark mass. It is seen from eq.(11) that the evaluation of the difference of the amplitudes within the discussed approach is mostly sensitive to the difference of the decay rates of $`\mathrm{\Xi }_c^0`$ and $`\mathrm{\Lambda }_c`$, with only very little sensitivity to the total decay width of $`\mathrm{\Xi }_c^+`$. Using the current data on the total decay rates: $`\mathrm{\Gamma }(\mathrm{\Lambda }_c)=4.85\pm 0.28ps^1`$, $`\mathrm{\Gamma }(\mathrm{\Xi }_c^0)=10.2\pm 2ps^1`$, and the updated value $`\mathrm{\Gamma }(\mathrm{\Xi }_c^+)=3.0\pm 0.45ps^1`$, the difference $`\mathrm{\Delta }A_S`$ is estimated as $$\mathrm{\Delta }A_S=(5.4\pm 2)\times 10^7,$$ (12) with the uncertainty being dominated by the experimental error in the lifetime of $`\mathrm{\Xi }_c^0`$. An $`S`$ wave amplitude $`A_S`$ of the magnitude, given by the central value in eq.(12) would produce a decay rate $`\mathrm{\Gamma }(\mathrm{\Xi }_Q\mathrm{\Lambda }_Q\pi )=|A_S|^2p_\pi /(2\pi )0.9\times 10^{10}s^1`$, which result can also be written in a form of triangle inequality $$\sqrt{\mathrm{\Gamma }(\mathrm{\Xi }_b^{}\mathrm{\Lambda }_b\pi ^{})}+\sqrt{\mathrm{\Gamma }(\mathrm{\Xi }_c^0\mathrm{\Lambda }_c\pi ^{})}\sqrt{0.9\times 10^{10}s^1}.$$ (13) Although at present it is not possible to evaluate in a reasonably model independent way the matrix element in eq.(5) for the ‘spectator’ decay amplitude, the estimate (13) shows that at least some of the discussed pion transitions should go at the level of $`0.01ps^1`$. Also the numerical result for $`\mathrm{\Delta }A_S`$ invites one more remark, related to the problem of the differences of lifetimes of heavy hadrons. Namely, we have seen here that the values of the matrix elements $`x`$ and $`y`$ extracted from the data on the lifetimes of charmed hyperons, translate, in terms of the pion transitions, into a ‘natural’ magnitude of the decay amplitude, which one would expect, based on the experience with the decays of the ordinary strange hyperons. On the other hand, these phenomenological numerical values of $`x`$ and $`y`$ are deemed to be substantially enhanced with respect to the simple estimates , $`y=x=f_D^2m_D/120.006GeV^3`$, based on non-relativistic-like ideas about the structure of the heavy baryons and mesons. The discussed here relation of these matrix elements to the decays $`\mathrm{\Xi }_Q\mathrm{\Lambda }_Q\pi `$ and the evaluation of their significance, perhaps, tell us that the phenomenological values of $`x`$ and $`y`$ are of a ‘normal’ magnitude, while the early simplistic theoretical estimates were simply too low. To summarize. It is shown that the strangeness-changing decays $`\mathrm{\Xi }_Q\mathrm{\Lambda }_Q\pi `$ of the $`b`$ and $`c`$ cascade hyperons should go at the rate of order $`0.01ps^1`$ and that they should obey the $`\mathrm{\Delta }I=1/2`$ rule. The decays $`\mathrm{\Xi }_b\mathrm{\Lambda }_b\pi `$ are purely $`S`$ wave in the heavy quark limit and their amplitude, as well as the $`S`$ wave amplitude of the decays $`\mathrm{\Xi }_s\mathrm{\Lambda }_s\pi `$, is expressed by the current algebra in terms of matrix elements of four-quark operators over the heavy baryons. The difference between the $`S`$ wave amplitude of the charmed hyperon decay and that of the $`\mathrm{\Xi }_b`$ is related, within the heavy quark expansion, to the differences of lifetimes among the heavy hyperons. The semileptonic transitions $`\mathrm{\Xi }_Q\mathrm{\Lambda }_Q\mathrm{}\nu `$ are of a purely $`0^+0^+`$ type and thus have $`g_A=0`$, $`g_V=1`$, while the radiative weak decay $`\mathrm{\Xi }_b\mathrm{\Lambda }_b\gamma `$ is shown to be greatly suppressed in the heavy quark limit. Acknowledgement This work is supported in part by DOE under the grant number DE-FG02-94ER40823.
warning/0001/math0001114.html
ar5iv
text
# Fermionic formulas for level-restricted generalized Kostka polynomials and coset branching functions ## 1. Introduction Generalized Kostka polynomials are $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}`$-analogues of the tensor product multiplicity (1.1) $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{dim}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{Hom}}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ where $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}`$ is a partition, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ is a sequence of rectangles and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}`$ is the irreducible integrable highest weight module of highest weight $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}`$ over the quantized enveloping algebra $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. The generalized Kostka polynomials can be expressed as generating functions of classically restricted paths . In terms of the theory of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$-crystals these paths correspond to the highest weight vectors of tensor products of perfect crystals. The statistic is given by the energy function on paths. The $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$-crystal structure can be extended to a $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$-crystal structure . For particular weights, the highest weight vectors of the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$-modules correspond to level-restricted paths. Hence it is natural to consider the generating functions of level-restricted paths, giving rise to level-restricted generalized Kostka polynomials which will take a lead rôle in this paper. The notion of level-restriction is also very important in the context of restricted-solid-on-solid (RSOS) models in statistical mechanics and fusion models in conformal field theory . The one-dimensional configuration sums of RSOS models are generating functions of level-restricted paths (see for example ). The structure constants of the fusion algebras of Wess–Zumino–Witten conformal field theories are exactly the level-restricted analogues of the Littlewood–Richardson coefficients in (1.1) as shown by Kac \[15, Exercise 13.35\] and Walton . $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}`$-Analogues of these level-restricted Littlewood–Richardson coefficients in terms of ribbon tableaux were proposed in ref. . The generalized Kostka polynomial admits a fermionic (or quasi-particle) formula . Fermionic formulas originate from the Bethe Ansatz which is a technique to construct eigenvectors and eigenvalues of row-to-row transfer matrices of statistical mechanical models. Under certain assumptions (the string hypothesis) it is possible to count the solutions of the Bethe equations resulting in fermionic expressions which look like sums of products of binomial coefficients. The Kostka numbers arise in the study of the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{X}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{X}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{X}$}`$ model in this way . Fermionic formulas are of interest in physics since they reflect the particle structure of the underlying model and also reveal information about the exclusion statistics of the particles . The fermionic formula of the Kostka polynomial can be combinatorialized by taking a weighted sum over sets of rigged configurations . In ref. the fermionic formula for the generalized Kostka polynomial was proven by establishing a statistic-preserving bijection between Littlewood–Richardson tableaux and rigged configurations. In this paper we show that this bijection is well-behaved with respect to level-restriction and we give an explicit characterization of level-restricted rigged configurations (see Definition 5.5 and Theorem 8.2). This enables us to obtain a combinatorial formula for the level-restricted generalized Kostka polynomials as the generating function of level-restricted rigged configurations (see Theorem 5.7). As an immediate consequence this proves a new general fermionic formula for the level-restricted generalized Kostka polynomial (see Theorem 6.2 and Eq. (6.7)). Special cases of this formula were conjectured in refs. . As opposed to some definitions of “fermionic formulas” the expression of Theorem 6.2 involves in general explicit negative signs. However, we would like to point out that because of the equivalent combinatorial formulation in terms of rigged configurations as given in Theorem 5.7 the fermionic sum is manifestly positive (i.e., a polynomial with positive coefficients). The branching functions of type $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}`$ can be described in terms of crystal graphs of irreducible integrable highest weight $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$-modules. For certain triples of weights they can be expressed as limits of level-restricted generalized Kostka polynomials. The structure of the rigged configurations allows one to take this limit, thereby yielding a fermionic formula for the corresponding branching functions (see Eq. (7.10)). The derivation of this formula requires the knowledge of the ground state energy, which is obtained from the explicit construction of certain local isomorphisms of perfect crystals (see Theorem 7.3). A more complete set of branching functions can be obtained by considering “skew” level-restricted generalized Kostka polynomials. We conjecture that rigged configurations are also well-behaved with respect to skew shapes (see Conjecture 8.3). The paper is structured as follows. Section 2 sets out notation used in the paper. In Section 3 we review some crystal theory, in particular the definition of level-restricted paths, which are used to define the level-restricted generalized Kostka polynomials. Littlewood–Richardson tableaux and their level-restricted counterparts are defined in Section 4. The formulation of the generalized Kostka polynomials in terms of Littlewood–Richardson tableaux with charge statistic is necessary for the proof of the fermionic formula which makes use of the bijection between Littlewood–Richardson tableaux and rigged configurations. The latter are subject of Section 5 which also contains the new definition of level-restricted rigged configurations and our main Theorem 5.7. The proof of this theorem is reserved for Section 8. The fermionic formulas for the level-restricted Kostka polynomial and the type $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}`$ branching functions are given in Sections 6 and 7, respectively. ### Acknowledgements We are deeply indebted to Anatol Kirillov for stimulating discussions. We would also like to thank Peter Bouwknegt, Srinandan Dasmahapatra, Atsuo Kuniba, and Masato Okado for helpful discussions. ## 2. Notation All partitions are assumed to have $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ parts, some of which may be zero. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ be a sequence of partitions whose Ferrers diagrams are rectangles. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$ have $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$ columns and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\eta }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$ rows for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}`$. We adopt the English notation for partitions and tableaux. Unless otherwise specified, all tableaux are assumed to be column-strict (that is, the entries in each row weakly increase from left to right and in each column strictly increase from top to bottom). ## 3. Paths The main goal of this section is to define the level-restricted generalized Kostka polynomials. These polynomials are defined in terms of certain finite $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$-crystal graphs whose elements are called paths. The theory of crystal graphs was invented by Kashiwara , who showed that the quantized universal enveloping algebras of Kac-Moody algebras and their integrable highest weight modules admit special bases whose structure at $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ is specified by a colored graph known as the crystal graph. The crystal graphs for the finite-dimensional irreducible modules for the classical Lie algebras were computed explicitly by Kashiwara and Nakashima . The theory of perfect crystals gave a realization of the crystal graphs of the irreducible integrable highest weight modules for affine Kac-Moody algebras, as certain eventually periodic sequences of elements taken from finite crystal graphs . This realization is used for the main application, some new explicit formulas for coset branching functions of type $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}`$. ### 3.1. Crystal graphs Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝔤}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ be the quantized universal enveloping algebra for the Kac-Moody algebra $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝔤}$}`$. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{I}$}`$ be an indexing set for the Dynkin diagram of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝔤}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}`$ the weight lattice of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝔤}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ the dual lattice, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{I}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}`$ the (not necessarily linearly independent) simple roots, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{I}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}`$ the simple coroots, and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{I}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}`$ the fundamental weights. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ denote the natural pairing of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}`$. Suppose $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}`$ is a $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝔤}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$-module with crystal graph $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$. Then $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ is a directed graph whose vertex set (also denoted $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$) indexes a basis of weight vectors of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}`$, and has directed edges colored by the elements of the set $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{I}$}`$. The edges may be viewed as a combinatorial version of the action of Chevalley generators. This graph has the property that for every $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{I}$}`$, there is at most one edge colored $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$ entering (resp. leaving) $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$. If there is an edge $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ colored $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$, denote this by $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$. If there is no edge colored $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$ leaving $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$ (resp. entering $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$) then say that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ (resp. $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$) is undefined. The $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$ are called Kashiwara lowering and raising operators. Define $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ (resp. $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$) to be the maximum $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ such that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ (resp. $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$) is defined. There is a weight function $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{wt}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}`$ that satisfies the following properties: (3.1) $$\begin{array}{cc}\hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{wt}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{wt}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\hfill \\ \hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{wt}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{wt}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\hfill \\ \hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{wt}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}\hfill \end{array}$$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ is called a $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}`$-weighted $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{I}$}`$-crystal. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{I}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}`$ be the set of dominant integral weights. For $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}`$ denote by $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝕍}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ the irreducible integrable highest weight $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝔤}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$-module of highest weight $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}`$. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝔹}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ be its crystal graph. Say that an element $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ of the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}`$-weighted $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{I}$}`$-crystal $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ is a highest weight vector if $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ for all $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{I}$}`$. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}`$ be the highest weight vector in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝔹}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. By (3.1), for all $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{I}$}`$, (3.2) $$\begin{array}{cc}\hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\hfill \\ \hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}\hfill \end{array}$$ Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ be the crystal graph of a $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝔤}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$-module $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$. A morphism of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}`$-weighted $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{I}$}`$-crystals is a map $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ such that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{wt}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{wt}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ for all $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{I}$}`$. In particular $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ is defined if and only if $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ is. Suppose $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ are $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝔤}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$-modules with crystal graphs $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ respectively. Then $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ admits a crystal graph denoted $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ which is equal to the direct product $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\times }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ as a set. We use the opposite of the convention used in the literature. Define (3.3) $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\begin{array}{cc}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\hfill & \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\text{if }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\text{,}}$}\hfill \\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\hfill & \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\text{if }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\text{ and }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\text{,}}$}\hfill \\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\text{undefined}}$}\hfill & \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\text{otherwise.}}$}\hfill \end{array}$$ Equivalently, (3.4) $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\begin{array}{cc}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\hfill & \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\text{if }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\text{,}}$}\hfill \\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\hfill & \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\text{if }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\text{ and }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\text{,}}$}\hfill \\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\text{undefined}}$}\hfill & \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\text{otherwise.}}$}\hfill \end{array}$$ One has (3.5) $$\begin{array}{cc}\hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{max}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\hfill \\ \hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{max}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}\hfill \end{array}$$ Finally $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{wt}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}`$ is defined by $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{wt}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{wt}}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{wt}}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ where $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{wt}}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{wt}}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}`$ are the weight functions for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$. This construction is ”associative”, that is, the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}`$-weighted $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{I}$}`$-crystals form a tensor category. ###### Remark 3.1. It follows from (3.4) that if $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ is defined, then $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ for some $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}`$. ### 3.2. $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$-crystal graphs on tableaux Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}`$ be the indexing set for the Dynkin diagram of type $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$, with weight lattice $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{fin}}$}}`$, simple roots $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}`$, fundamental weights $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}`$, and simple coroots $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}`$. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ be a partition. There is a natural projection $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{fin}}$}}`$ denoted $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ be the irreducible integrable highest weight module of highest weight $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}`$ over the quantized universal enveloping algebra $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ . By abuse of notation we shall write $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ and denote the crystal graph of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}`$ by $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}`$. As a set $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}`$ may be realized as the set of tableaux of shape $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}`$ over the alphabet $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}`$. Define the content of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}`$ by $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{content}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ where $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$ is the number of times the letter $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$ appears in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$. The weight function $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{wt}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{fin}}$}}`$ is given by sending $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$ to the image of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{content}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ under the projection $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{fin}}$}}`$. The row-reading word of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$ is defined by $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{word}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ where $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}`$ is the word obtained by reading the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}`$-th row of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$ from left to right. This definition is useful even in the context that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$ is a skew tableau. The edges of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}`$ are given as follows. First let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}`$ be a word in the alphabet $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}`$. View each letter $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$ (resp. $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$) of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}`$ as a closing (resp. opening) parenthesis, ignoring other letters. Now iterate the following step: declare each adjacent pair of matched parentheses to be invisible. Repeat this until there are no matching pairs of visible parentheses. At the end the result must be a sequence of closing parentheses (say $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}`$ of them) followed by a sequence of opening parentheses (say $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}`$ of them). The unmatched (visible) subword is of the form $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}`$. If $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ (resp. $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$) then $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ (resp. $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$) is obtained from $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}`$ by replacing the unmatched subword $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}`$ by $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ (resp. $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$). Then $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}`$, and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ (resp. $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$) is defined if and only if $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ (resp. $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$). For the tableau $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}`$, let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ be undefined if $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{word}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ is; otherwise define $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ to be the unique (not necessarily column-strict) tableau of shape $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}`$ such that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{word}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{word}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. It is easy to verify that when defined, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ is a column-strict tableau. Consequently $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{word}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. The operator $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$ and the quantity $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ are defined similarly. ### 3.3. $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$-crystal structure on rectangular tableaux There is an inclusion of algebras $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ where $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ is the quantized universal enveloping algebra corresponding to the derived subalgebra $`\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ of the affine Kac-Moody algebra $`\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ . Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{I}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}`$ be the index set for the Dynkin diagram of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cl}}$}}`$ be the weight lattice of $`\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$, with (linearly dependent) simple roots $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cl}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{I}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}`$, simple coroots $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{I}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}`$, and fundamental weights $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cl}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{I}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}`$. The simple roots satisfy the relation $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cl}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cl}}$}}`$. There is a natural projection $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cl}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{fin}}$}}`$ with kernel $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ such that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cl}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cl}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cl}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{fin}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cl}}$}}`$ be the section of the above projection defined by $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cl}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cl}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cl}}$}}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}`$. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ be the canonical central element. The level of a weight $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cl}}$}}`$ is defined by $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cl}}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cl}}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}`$. Suppose $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}`$ is a finite-dimensional $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$-module that has a crystal graph $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ (not all do); $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ is a $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cl}}$}}`$-weighted $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{I}$}`$-crystal. A weight function $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{wt}}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cl}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cl}}$}}`$ may be given by $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{wt}}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cl}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cl}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{wt}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ where $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{wt}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{fin}}$}}`$ is the weight function on the set $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ viewed as a $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$-crystal graph. In addition to being a $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$-crystal graph, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ also has some edges colored $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$. The action of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ on $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}`$ extends to an action of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ which admits a crystal structure, if and only if the partition $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}`$ is a rectangle . If $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}`$ is the rectangle with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}`$ rows and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}`$ columns, then write $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}}`$ for the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$-module with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$-structure $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}`$ and denote its crystal graph by $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}}`$. If one of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}`$ or $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}`$ is $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$, then it is easy to give $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ explicitly on $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}}`$, for in this case the weight spaces of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}}`$ are one-dimensional, and the zero edges can be deduced from (3.1) . The general case is given as follows . We shall first define a content-rotating bijection $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}}`$. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}}`$ be a tableau, say of content $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ will have content $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Remove all the letters $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ from $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$, leaving a vacant horizontal strip of size $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ in the northwest corner of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$. Compute Schensted’s $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}`$ tableau of the row-reading word of this skew subtableau. It can be shown that this yields a tableau of the shape obtained by removing $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ cells from the last row of the rectangle $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Subtract one from the value of each entry of this tableau, and then fill in the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ vacant cells in the last row of the rectangle $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ with the letter $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$. It can be shown that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ is a well-defined bijection, whose inverse $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}`$ can be given by a similar algorithm. Then (3.6) $$\begin{array}{cc}\hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\hfill \\ \hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}\hfill \end{array}$$ for all $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$ where indices are taken modulo $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$; in particular for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ this defines explicitly the operators $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$. ### 3.4. Sequences of rectangular tableaux For a sequence of rectangles $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$, consider the tensor product $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$. Its $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$-crystal graph has underlying set $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒫}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$, where the tensor symbols denote the Cartesian product of sets. A typical element of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒫}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$ is called a path and is written $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ where $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}}`$ is a tableau of shape $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$. The edges of the crystal graph $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒫}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$ are given explicitly as follows. Define the word of a path $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$ by $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{word}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{word}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{word}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{word}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ Then for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ (as in the definition of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}`$), if $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{word}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ is undefined, let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ be undefined; otherwise it not hard to see that there is a unique path $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒫}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$ such that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{word}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{word}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. To define $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$, let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}`$. This definition is equivalent to that given by taking the above definition of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$ on the crystals $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}}`$ and then applying the rule for lowering operators on tensor products (3.3). The action of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{I}$}`$ is defined analogously. ### 3.5. Integrable affine crystals Consider the affine Kac-Moody algebra $`\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$, with weight lattice $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{af}}$}}`$, independent simple roots $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{I}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}`$, simple coroots $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{I}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}`$, and fundamental weights $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{I}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}`$. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{af}}$}}`$ be the null root. There is a natural projection which we shall by abuse of notation also call $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cl}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{af}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cl}}$}}`$ such that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cl}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cl}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cl}}$}}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{I}$}`$. Write $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{af}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cl}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{af}}$}}`$ for the section of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cl}}$}`$ given by $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{af}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cl}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{I}$}`$. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cl}}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}`$ be a dominant integral weight and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝔹}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ the crystal graph of the irreducible integrable highest weight $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$-module of highest weight $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}`$. If $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ then $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝔹}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ is infinite. The set of weights in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{af}}$}}`$ that project by $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cl}}$}`$ to $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}`$ are given by $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cl}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{af}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}`$. Now fix $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$. The irreducible integrable highest weight $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$-crystal graph $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝔹}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{af}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ may be identified with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝔹}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ as sets and as $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{I}$}`$-crystals (independent of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$). The weight functions for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝔹}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{af}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝔹}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{af}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ differ by the global constant $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}`$. The weight function $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝔹}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ is obtained by composing the weight function for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝔹}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{af}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$, with the projection $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cl}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{af}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cl}}$}}`$. The set $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝔹}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ is then endowed with an induced $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$-grading $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝔹}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ defined by $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{wt}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ where $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝔹}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ is identified with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝔹}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{af}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{wt}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝔹}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{af}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{af}}$}}`$ is the weight function and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{af}}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ is the degree generator. The map $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ takes the coefficient of the element $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}`$ of an element in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{af}}$}}`$ when written in the basis $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{I}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}`$. ### 3.6. Energy function on finite paths The set of paths $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒫}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$ has a natural statistic called the energy function. The definitions here follow . Consider first the case that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ is a sequence of two rectangles. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$. Since $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ is a connected crystal graph, there is a unique $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$-crystal graph isomorphism (3.7) $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ This is called the local isomorphism (see Section 4.4 for an explicit construction). Write $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$. Then there is a unique (up to a global additive constant) map $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ such that (3.8) $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\begin{array}{cc}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\hfill & \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\text{if }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\text{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\hfill \\ & \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\text{and }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\text{,}}$}\hfill \\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\hfill & \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\text{if }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\text{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\hfill \\ & \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\text{and }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\text{,}}$}\hfill \\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\hfill & \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\text{otherwise.}}$}\hfill \end{array}$$ This map is called the local energy function. By definition it is invariant under the local isomorphism and under $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}`$. Let us normalize it by the condition that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}`$ where $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$ is the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ highest weight vector of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$ is the intersection of the Ferrers diagrams of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$, and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}`$ is the number of cells in this intersection. Explicitly $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{min}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\eta }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\eta }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{min}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}`$. If $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\eta }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\eta }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ then the local energy function attains precisely the values from $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ to $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}`$. Now let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ be a sequence of rectangles and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒫}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$. For $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}`$ denote the local isomorphism that exchanges the tensor factors in the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}`$-th and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$-th positions. For $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}`$, let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ be the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$-th tensor factor in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Then define the energy function (3.9) $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ The value of the energy function is unchanged under local isomorphisms and under $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}`$, since the local energy function has this property. The next lemma follows from the definition of the local energy function. ###### Lemma 3.2. Suppose $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒫}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$ is such that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ is defined and for any image $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$ under a composition of local isomorphisms, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ where $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. Then $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. If all rectangles $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$ are the same then each of the local isomorphisms is the identity and (3.10) $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ Say that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒫}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$ is classically restricted if it is an $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$-highest weight vector, that is, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ for all $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}`$. Equivalently, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{word}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ is a (reverse) lattice permutation (every final subword has partition content). Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒫}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}}`$ be the set of classically restricted paths in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒫}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$ of weight $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cl}}$}}`$. It was shown in that the generalized Kostka polynomial (which was originally defined in terms of Littlewood–Richardson tableaux; see (4.3)) can be expressed as (3.11) $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{K}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒫}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cl}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ This extends the path formulation of the Kostka polynomial by Nakayashiki and Yamada . ### 3.7. Level-restricted paths Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ be any $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cl}}$}}`$-weighted $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{I}$}`$-crystal and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cl}}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}`$. Say that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ is $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}`$-restricted if $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}`$ is a highest weight vector in the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cl}}$}}`$-weighted $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{I}$}`$-crystal $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝔹}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$, that is, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ for all $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{I}$}`$. Equivalently $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ for all $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{I}$}`$ by (3.5) and (3.2). Denote by $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ the set of elements $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ that are $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}`$-restricted. If $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cl}}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}`$ has the same level as $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}`$, define $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ to be the set of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ such that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{wt}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cl}}$}}`$, that is, the set of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ such that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}`$ is a highest weight vector of weight $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$. Say that the element $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$ is restricted of level $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$ if it is $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$-restricted. Such paths are also classically restricted since $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}`$. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒫}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}`$ denote the set of paths in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒫}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}}`$ that are restricted of level $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$. Letting $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒫}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$, this is the same as saying $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒫}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Define the level-restricted generalized Kostka polynomial by (3.12) $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{K}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒫}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cl}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ ### 3.8. Perfect crystals This section is needed to compute the coset branching functions in Section 7. We follow , stating the definitions in the case of $`\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$. For any $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$-crystal $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$, define $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cl}}$}}`$ by $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{I}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{I}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$. Now let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$ be a positive integer and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ the crystal graph of a finite dimensional irreducible $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$-module $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}`$. Say that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ is perfect of level $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$ if 1. $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ is connected. 2. There is a weight $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cl}}$}}`$ such that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ has a unique vector of weight $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ and all other vectors in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ have lower weight in the Chevalley order, that is, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{wt}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$. 3. $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{min}}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$. 4. The maps $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}`$ restrict to bijections $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{min}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cl}}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}`$ where $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{min}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ is the set of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ achieving the minimum in 3. For $`\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ the perfect crystals of level $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$ are precisely those of the form $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ . Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}`$. The weight $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ can be taken to be $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cl}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cl}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. ###### Example 3.3. We describe the bijections $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{min}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cl}}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}`$ in this example. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}`$. For this example let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{6}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}`$, and consider the weight $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}`$. As usual subscripts are identified modulo $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$. The unique tableau $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}`$ such that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}`$ is constructed as follows. First let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}`$ be the following tableau of shape $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Its bottom row contains $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ copies of the letter $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ (here it is $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{12466}$}`$ since the sequence of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{6}$}`$ is $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$). Let every letter in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}`$ have value one smaller than the letter directly below it. Here we have $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\begin{array}{ccccc}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{6}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{6}$}\end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ be the subtableau of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}`$ consisting of the entries that are nonpositive and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}`$ the rest. Say $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ has shape $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}`$ (here $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$). Let $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ (here $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$). The desired tableau $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$ is defined as follows. The restriction of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$ to the shape $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}`$ is $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$, or equivalently, the tableau obtained by taking the skew tableau $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}`$ and first pushing all letters straight upwards to the top of the bounding rectangle $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$, and then pushing all letters straight to the left inside $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. The restriction of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$ to $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}`$ is the tableau of that skew shape in the alphabet $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}`$ with maximal entries, that is, its bottom row is filled with the letter $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$, the next-to-bottom row is filled with the letter $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$, etc. In the example, $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\begin{array}{ccccc}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{6}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{6}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{6}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{6}$}\end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ To construct the unique element $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}`$ such that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}`$, let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}`$ be the tableau whose first row has $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ copies of the letter $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$, again identifying subscripts modulo $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$; here $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}`$ has first row $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{11235}$}`$. Now let the rest of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}`$ be defined by letting each entry have value one greater than the entry above it. So $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\begin{array}{ccccc}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{6}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{7}$}\end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ be the subtableau of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}`$ consisting of the values that are at most $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}`$ be the shape of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ and $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Here $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ and $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. The element $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ is defined as follows. Its restriction to the skew shape $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}}`$ is the unique skew tableau $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}`$ of that shape such that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$, or equivalently, this restriction is obtained by taking the tableau $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$, pushing all letters directly down within the rectangle $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ and then pushing all letters to the right within $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. The restriction of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ to the shape $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}}`$ is filled with the smallest letters possible, so that the first row of this subtableau consists of ones, the second row consists of twos, etc. Here $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\begin{array}{ccccc}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{6}$}\end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ The main theorem for perfect crystals is: ###### Theorem 3.4. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ be a perfect crystal of level $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cl}}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}`$ with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$. Then there is an isomorphism of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$-crystals (3.13) $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝔹}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝔹}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{wt}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ Suppose now that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ is perfect of level $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cl}}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}`$. Write $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ for the unique element of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ such that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}`$. Theorem 3.4 (with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}`$ therein replaced by $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$) says that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝔹}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝔹}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ with corresponding highest weight vectors $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}`$. This isomorphism can be iterated. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{min}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{min}}$}}`$ be the unique bijection defined by $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}`$. Then there are isomorphisms $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝔹}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝔹}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ such that the highest weight vector of the left-hand side is given by $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$. For the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ perfect crystals $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}`$, it can be shown that the map $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}`$ is none other than the power $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}`$ of the content rotating map $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}`$. Moreover if $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}`$ is extended to a bijection $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}`$ by defining $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}`$, then the extended function also satisfies $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ for all $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}`$ not just for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{min}}$}}`$. Since the bijection $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}`$ on $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}`$ has order $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$, the bijection $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}`$ has order $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{gcd}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. The ground state path for the pair $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ is by definition the infinite periodic sequence $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$ where $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ be the set of all semi-infinite sequences $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$ of elements in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ such that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$ eventually agrees with the ground state path $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Then the set $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ has the structure of the crystal $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝔹}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ with highest weight vector $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}`$ and weight function $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{wt}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{wt}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{wt}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. To recover the weight function of the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$-crystal $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝔹}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{af}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$, define the energy function on $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ by (3.14) $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}$$ and define the map $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝔹}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{af}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{af}}$}}`$ by $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{wt}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}`$ where $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{wt}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝔹}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cl}}$}}`$. $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ can be regarded as a direct limit of the finite crystals $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}`$. Define the embedding $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ by $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}$$ Define $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ by $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ where the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}`$ on the right hand side is the energy function for the finite path space $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$. By definition for all $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Note that the last fixed step $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ is necessary to make the energy function on the finite paths stable under the embeddings into $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. ### 3.9. Standardization embeddings We require certain embeddings of finite path spaces. Given a sequence of rectangles $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$, let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{r}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ denote the sequence of rectangles given by splitting the rectangles of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$ into their constituent rows. For example, if $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ then $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{r}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. There is a unique embedding (3.15) $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒫}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒫}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{r}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}$$ defined as follows. Its explicit computation is based on transforming $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$ into $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{r}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ using two kinds of steps. 1. Suppose $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ has more than one row ($`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\eta }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$). Then use the transformation $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\eta }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Informally, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}`$ is obtained from $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$ by splitting off the first row of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. There is an associated embedding of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$-crystal graphs $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒫}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒫}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}}`$ defined by the property that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{word}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{word}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ for all $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒫}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$. Here it is crucial that the rectangle being split horizontally, is the first one, for otherwise the embedding does not preserve the edges labeled by $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$. 2. If $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\eta }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$, then use a transformation of the form $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$ for some $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}`$. Here $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$ denotes the sequence of rectangles obtained by exchanging the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}`$-th and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$-th rectangles in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$. The associated isomorphism of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$-crystal graphs is the local isomorphism $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒫}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒫}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}}`$ defined before. It is clear that one can transform $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$ into $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{r}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ using these two kinds of steps. Now fix one such sequence of steps leading from $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$ to $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{r}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$, say $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{r}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ where each $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ is a sequence of rectangles and each step $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ is one of the two types defined above. Define the map $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒫}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒫}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}}`$ by $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}`$ if the step is of the first kind, and by $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}`$ if it is of the second kind. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒫}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒫}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{r}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ be the composition $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$. It can be shown that the map $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$ does not depend on the sequence of the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$; this is proven in the equivalent language of Littlewood–Richardson tableaux in . ## 4. Littlewood–Richardson tableaux We now review some formulations of type $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}`$ tensor product multiplicities that use tableaux. These tableaux, which we call Littlewood–Richardson (LR) tableaux, are the intermediate combinatorial objects between paths and rigged configurations, which give rise to fermionic expressions. For the most part, the material in this section is taken from . ### 4.1. Three formulations Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{I}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{I}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{I}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}`$ be intervals of integers such that if $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{I}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{I}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$, then $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}`$. Set $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{I}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{I}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$. For each $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}`$, fix a tableau $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$ of shape $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$ in the alphabet $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{I}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$. Define the set $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{SLR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ to be the set of tableaux $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}`$ of shape $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}`$ in the alphabet $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{I}$}`$ such that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{I}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$ for all $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$, where $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{I}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}}`$ denotes the skew subtableau of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}`$ obtained by restricting to the alphabet $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{I}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$, and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ denotes the Schensted $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}`$-tableau of the row-reading word of the skew tableau $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}`$. It is well-known that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{SLR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}`$, where $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}`$ was defined in (1.1). We shall define three kinds of LR tableaux given by $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{SLR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ for various choices of intervals $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{I}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$ and tableaux $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$. 1. $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{LR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$: Define the set of intervals of integers $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{I}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{[}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\eta }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\eta }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\eta }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\eta }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\eta }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{]}$}`$. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$ be the tableau of shape $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$ whose $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}`$-th row is filled with copies of the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}`$-th largest letter of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$, namely, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\eta }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\eta }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}`$. Define $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{LR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{SLR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. When $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$ consists of single rows (that is, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\eta }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ for all $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$), then $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{LR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{CST}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$, the (column-strict) tableaux of shape $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}`$ and content $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}`$. 2. $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{CLR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ (Columnwise LR): Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ be the standard tableau of shape $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ obtained by placing the numbers $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ through $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\eta }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ down the first column, the next $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\eta }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ numbers down the second column, etc. Continue this process to obtain $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$, starting with the next available number, namely, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\eta }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. Explicitly, for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}`$, the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$-th entry in the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$-th tableau $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$ is equal to $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\eta }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\eta }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\eta }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}`$. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{B}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$ be the interval consisting of the entries of the tableau $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$. Define $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{CLR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{SLR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. 3. $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RLR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ (Rowwise LR): Define this similarly to $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{CLR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ but label by rows, so that the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$-th entry of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$ is $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\eta }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\eta }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}`$. Then let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RLR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{SLR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. ###### Example 4.1. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Here $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}`$, and $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\begin{array}{c}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\text{and}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\begin{array}{cc}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ We have $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{B}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{B}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}`$, $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\begin{array}{c}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\text{and}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\begin{array}{cc}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}\end{array}$$ and $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\begin{array}{c}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\text{and}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\begin{array}{cc}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}\end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ Observe that $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\begin{array}{ccc}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}& \end{array}$$ is in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{CLR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ since $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{B}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{B}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$. On the other hand $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\begin{array}{ccc}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\end{array}`$ is not in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{CLR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ since $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{B}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\begin{array}{ccc}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}& & \end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$. ### 4.2. Obvious bijections among the various LR tableaux There are trivial relabeling bijections between the various kinds of LR tableaux defined above. We give them explicitly here for later use. 1. The bijection $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{CLR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RLR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ is given by the following relabeling. Consider an entry $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}`$ in a standard tableau $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{CLR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Then $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}`$ appears in one of the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}`$ tableaux, say, it is the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$-th entry of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}`$ be the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$-th entry of the rowwise tableau $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$. Then replace $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}`$ by $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}`$ in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}`$. Performing all such replacements simultaneously yields $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RLR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. 2. The bijection $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{std}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{LR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RLR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ is given by Schensted’s standardization map . Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{LR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$ be some entry in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}`$. Suppose $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$ is the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}`$-th largest value in the subinterval $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$. Replace the occurrences of the letter $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$ in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}`$ from left to right by the consecutive integers given by the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}`$-th row of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$. The result of these substitutions is $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{std}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RLR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. 3. Define a bijection $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{LR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{CLR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ by $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{std}}$}`$. 4. Observe that ordinary transposition of standard tableaux restricts to a bijection $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tr}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RLR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{CLR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ where $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}`$ denotes the transpose partition of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. 5. There is a bijection $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tr}}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{LR}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{CLR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{CLR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ defined by $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tr}}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{LR}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tr}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$. ### 4.3. Paths to tableau pairs The Robinson–Schensted–Knuth correspondence allows one to pass from paths to pairs of tableaux. This bijection gives a combinatorial decomposition of the crystal graph of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒫}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$ into $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ irreducible components and encodes the energy function in the recording tableau. The column insertion version of the Robinson–Schensted–Knuth correspondence, restricts to a bijection (4.1) $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RSK}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒫}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{CST}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\times }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{LR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}$$ as follows. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒫}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$. Define $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{word}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. This can be computed by the column insertion of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{word}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ starting from the right end. Recall that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$ are column-strict tableaux of shape $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ be the tableau obtained by recording the insertion of a letter in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$ by the letter in the corresponding position in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$. It can be shown that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{LR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$, and that the map (4.1) given by $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ is a bijection. ###### Remark 4.2. 1. This bijection is a morphism of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$-crystal graphs in the sense that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}`$. In particular, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒫}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$ is classically restricted if and only if $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ is a Yamanouchi tableau, that is, its $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}`$-th row is filled with copies of the letter $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}`$ for all $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$. 2. The energy function on paths can be transferred easily to a statistic on $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{LR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ called the generalized charge (written $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$) such that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. The generalized charge is defined explicitly in (4.2) below. ###### Example 4.3. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒫}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$ given by $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\begin{array}{cc}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\begin{array}{c}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ Then $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{word}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{2211\hspace{0.17em}1}}$}`$ and $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\begin{array}{ccc}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}& \end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\begin{array}{ccc}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}& \end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ ### 4.4. Generalized Automorphisms of Conjugation For the moment let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}`$ Recall that the local isomorphism (3.7) is the unique isomorphism of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$-crystal graphs $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$ or equivalently $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒫}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒫}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$. Let us make this more explicit. By Remark 4.2 we have a commutative diagram of bijections $$\begin{array}{ccc}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒫}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}& \stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RSK}}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{CST}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\times }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{LR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}& & \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\times }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}& & \\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒫}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}& \underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RSK}}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{CST}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\times }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{LR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\end{array}$$ such that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. This induces a bijection $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{LR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{LR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ for each $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}`$. The tensor product $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ is multiplicity-free. Therefore the domain and codomain of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}`$ are both empty or both singletons. Hence the bijection $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}`$ is unique and can be computed from the definition of the set $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{LR}}$}`$. Then $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ can be computed by applying RSK to obtain $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$, then applying $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}`$ to get $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$, and finally, the inverse of RSK to obtain $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. The local energy function is recovered using only the shape of the tableau pair. For a tableau $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{LR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ be the number of cells in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}`$ that lie strictly to the right of the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{max}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}`$-th column, or equivalently, strictly to the right of the shape $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$. Then $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Then the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$-crystal graph isomorphism $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒫}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒫}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}}`$ induces involutions $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{LR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{LR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ such that the diagram commutes: $$\begin{array}{ccc}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒫}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}& \stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RSK}}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{CST}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\times }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{LR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}& & \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\times }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}& & \\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒫}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}}& \underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RSK}}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{CST}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\times }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{LR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}\end{array}$$ The map $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}`$ is computed explicitly as follows . Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{LR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$ be the alphabets as in the definition of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{LR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Remove the skew subtableau $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}}`$. Use the usual column insertion of its row reading word, obtaining a pair of tableaux $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ where $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{LR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ for some partition $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ is the standard column insertion tableau. Next replace $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ by $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ where $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}`$ is the unique bijection $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{LR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{LR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Finally, pull back the pair of tableaux $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ under column insertion to obtain a word which turns out to be the row reading word of a skew column-strict tableau $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}`$ of the same shape as $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}`$. Then $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ is obtained by replacing $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}`$ by $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}`$. The bijections $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}`$ specialize to the automorphisms of conjugation of Lascoux and Schützenberger in the case that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$ consists of single rows. It is shown in that the bijections $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}`$ define an action of the symmetric group $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}`$ on paths and LR tableaux respectively. Specifically, for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}`$ let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}}`$ be any factorization of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}`$ into adjacent transpositions $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. For $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒫}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$, define $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒫}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}}`$. For $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{LR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ define $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{LR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. ### 4.5. Generalized charge The generalized charge on $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{LR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ is defined by (4.2) $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{!}$}}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ where $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{word}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ where $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}`$ is understood to be the function $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{LR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$. It was shown in \[33, Section 6\] and that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{LR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{LR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ has the structure of a graded poset with covering relation given by the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$-cocyclage and grading function given by the generalized charge. The generalized Kostka polynomial is by definition the generating function of LR tableaux with the charge statistic (4.3) $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{K}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{LR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ This extends the charge representation of the Kostka polynomial $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{K}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ of Lascoux and Schützenberger . For a path $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒫}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$ one has $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ , so the formulas (3.11) and (4.3) are equivalent. ### 4.6. Embeddings of LR tableaux The embeddings (3.15) of sets of paths, induce embeddings (4.4) $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{LR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{LR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{r}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}$$ via RSK. These maps are defined in . In the notation of \[25, Section 8.4\] they are denoted $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{r}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$. They are given by compositions of the generalized automorphisms of conjugation $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}`$ and by the embeddings of the form $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{LR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{LR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ (which is just the inclusion map). These embeddings preserve the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$-cocyclage poset structure and the generalized charge, since they are induced by maps that preserve the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$-crystal graph structure. ### 4.7. Level-restricted LR tableaux Say that a tableau $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{LR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ is restricted of level $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$ if there is a level-restricted path $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒫}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}`$ such that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Denote the set of such tableaux by $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{LR}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. ###### Example 4.4. Suppose each rectangle is a single row so that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{LR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{CST}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. In this case let us write $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{CST}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{LR}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. The following explicit rule appears in . Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{CST}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. The tableau $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}`$ may be viewed as a sequence of shapes $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}`$ where $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ is the shape of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{[}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{]}$}}`$. Then $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}`$ is restricted of level $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$ if (4.5) $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\text{ for all }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\text{.}}$}$$ In the further special case that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ for all $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$, write $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{ST}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{LR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ for the set of standard tableaux of shape $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}`$ and write $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{ST}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{LR}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ for the level-restricted subset. For $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{ST}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$, associate the chain of shapes $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ as above. Since passing from $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ to $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ adds only one additional cell, the condition (4.5) simplifies to (4.6) $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\text{ for all }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\text{.}}$}$$ For general $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$ it is possible to transfer the condition of level-restriction on paths to an explicit condition on LR tableaux. However for our purposes it is more convenient to use the following description of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{LR}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Since the embedding (4.4) is induced by the embedding (3.15) that preserves the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$-crystal graph structure, it follows that (4.7) $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{LR}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{LR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{CST}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{r}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ Hence an expression for the level-restricted generalized Kostka polynomials equivalent to (3.12) is $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{K}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{LR}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ ## 5. Rigged configurations This section follows \[25, Section 2.2\], with the notational difference that here $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$ is a rectangle with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$ columns and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\eta }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$ rows. The reason for this is that here we work with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RC}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ rather than $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RC}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ as in . ### 5.1. Review of definitions A $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$-configuration is a sequence of partitions $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ with the size constraints (5.1) $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{max}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\eta }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}$$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ where by convention $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ is the empty partition. If $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}`$ has at most $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ parts all partitions $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ are empty. For a partition $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}`$, define $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ to be the number of parts equal to $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$ and $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{min}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ the size of the first $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$ columns of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}`$. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\xi }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ be the partition whose parts are the widths of the rectangles in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$ of height $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}`$. The vacancy numbers for the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$-configuration $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}`$ are the numbers (indexed by $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$) defined by (5.2) $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\xi }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ In particular $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ for all $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. The $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$-configuration $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}`$ is said to be admissible if $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ for all $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$, and the set of admissible $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$-configurations is denoted by $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{C}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Following \[26, (3.2)\], set $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cc}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}$$ where $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ is the size of the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$-th column in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$. Define the charge $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ of a configuration $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{C}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ by $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cc}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}$$ $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\text{with}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\text{and}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ Observe that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ depends on both $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$ but $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cc}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ depends only on $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}`$. ###### Example 5.1. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Then $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ is a $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$-configuration with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\xi }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\xi }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. The configuration $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}`$ may be represented as $$\begin{array}{ccc}& & \\ \hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}& & \end{array}\begin{array}{ccc}& & \\ \hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}& & \\ & & \\ \hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}& & \end{array}\begin{array}{cc}& \\ \hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}& \end{array}$$ where the vacancy numbers are indicated to the left of each part. In addition $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cc}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. Define the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}`$-binomial by $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left[}$}\genfrac{}{}{0.0pt}{}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right]}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}}$$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ and zero otherwise where $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. The following fermionic or quasi-particle expression of the generalized Kostka polynomials, is a variant of \[25, Theorem 2.10\]. ###### Theorem 5.2. For $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}`$ a partition and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$ a sequence of rectangles (5.3) $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{K}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{C}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left[}$}\genfrac{}{}{0.0pt}{}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right]}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ Expression (5.3) can be reformulated as the generating function over rigged configurations. To this end we need to define certain labelings of the rows of the partitions in a configuration. For this purpose one should view a partition as a multiset of positive integers. A rigged partition is by definition a finite multiset of pairs $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ where $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$ is a positive integer and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}`$ is a nonnegative integer. The pairs $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ are referred to as strings; $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$ is referred to as the length of the string and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}`$ as the label or quantum number of the string. A rigged partition is said to be a rigging of the partition $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}`$ if the multiset consisting of the lengths of the strings, is the partition $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}`$. So a rigging of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}`$ is a labeling of the parts of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}`$ by nonnegative integers, where one identifies labelings that differ only by permuting labels among equal-sized parts of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}`$. A rigging $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}`$ of the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$-configuration $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}`$ is a sequence of riggings of the partitions $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ such that for every part of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ of length $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$ and label $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}`$, (5.4) $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ The pair $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ is called a rigged configuration. The set of riggings of admissible $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$-configurations is denoted by $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RC}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ be the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}`$-th rigged partition of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. A string $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ is said to be singular if $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$, that is, its label takes on the maximum value. Observe that the definition of the set $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RC}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ is completely insensitive to the order of the rectangles in the sequence $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$. However the notation involving the sequence $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$ is useful when discussing the bijection between LR tableaux and rigged configurations, since the ordering on $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$ is essential in the definition of LR tableaux. Define the cocharge and charge of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RC}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ by $$\begin{array}{cc}\hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cc}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cc}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\hfill \\ \hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\hfill \\ \hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\hfill \end{array}$$ where $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ is the partition inside the rectangle of height $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ and width $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ given by the labels of the parts of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ of size $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$. Since the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}`$-binomial $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left[}$}\genfrac{}{}{0.0pt}{}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right]}$}`$ is the generating function of partitions with at most $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}`$ parts each not exceeding $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}`$ \[1, Theorem 3.1\], Theorem 5.2 is equivalent to the following theorem. ###### Theorem 5.3. For $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}`$ a partition and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$ a sequence of rectangles (5.5) $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{K}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RC}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ ### 5.2. Switching between quantum and coquantum numbers Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RC}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RC}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ be the involution that complements quantum numbers. More precisely, for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RC}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$, replace every string $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ by $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. The notation here differs from that in , in which $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$ is an involution on $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RC}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. ###### Lemma 5.4. $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cc}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ for all $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RC}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. ###### Proof. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. It follows immediately from the definitions that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}`$. In particular $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ have the same vacancy numbers and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}`$. Then $$\begin{array}{cc}\hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cc}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\hfill \\ & \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cc}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cc}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}\hfill \end{array}$$ There is a bijection $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tr}}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RC}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RC}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RC}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ that has the property (5.6) $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cc}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tr}}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RC}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cc}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}$$ for all $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RC}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$; see the proof of \[26, Proposition 11\]. ### 5.3. RC’s and level-restriction Here we introduce the most important new definition in this paper, namely, that of a level-restricted rigged configuration. Say that a partition $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}`$ is restricted of level $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$ if $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$, recalling that it is assumed that all partitions have at most $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ parts, some of which may be zero. Fix a shape $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}`$ and a sequence of rectangles $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$ that are all restricted of level $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$. Define $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$, which is nonnegative by assumption. Set $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}`$ and denote the set of all column-strict tableaux of shape $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ over the alphabet $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}`$ by $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{CST}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Define a table of modified vacancy numbers depending on $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{C}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{CST}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ by (5.7) $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}$$ for all $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$, where $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ if the statement $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}`$ is true and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ otherwise, and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}`$ is the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$-th entry of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}`$. Finally let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ be the largest part of the partition $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$; if $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ is empty set $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$. ###### Definition 5.5. Say that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RC}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ is restricted of level $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$ provided that 1. $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$ for all $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}`$. 2. There exists a tableau $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{CST}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$, such that for every $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$, $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{C}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ be the set of all $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{C}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ such that the first condition holds, and denote by $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RC}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ the set of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RC}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ that are restricted of level $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$. Note in particular that the second condition requires that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ for all $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. ###### Example 5.6. Let us consider Definition 5.5 for two classes of shapes $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}`$ more closely: 1. Vacuum case: Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ be rectangular with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ rows. Then $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ for all $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ so that the modified vacancy numbers are equal to the vacancy numbers. 2. Two-corner case: Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$. Then $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ and there is only one tableau $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}`$ in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{CST}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$, namely the Yamanouchi tableau of shape $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$. Since $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}`$ we find that $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{max}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}$$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$. We wish to thank Anatol Kirillov for communicating this formula to us . Our main result is the following formula for the level-restricted generalized Kostka polynomial: ###### Theorem 5.7. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$ be a positive integer. For $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}`$ a partition and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$ a sequence of rectangles both restricted of level $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$, $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{K}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RC}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ The proof of this theorem is given in Section 8. ###### Example 5.8. Consider $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Then (5.8) $$\begin{array}{cc}& \\ \hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}& \\ & \\ \hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}& \\ & \\ \hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}& \end{array}\begin{array}{cc}& \\ \hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}& \end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\text{and}}$}\begin{array}{ccc}& & \\ \hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}& & \\ & & \\ \hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}& & \end{array}\begin{array}{cc}& \\ \hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}& \end{array}$$ are in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{C}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ where again the vacancy numbers are indicated to the left of each part. The set $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{CST}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ consists of the two elements $$\begin{array}{cc}& \\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\\ & \\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}& \end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\text{and}}$}\begin{array}{cc}& \\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\\ & \\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}& \end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ Since $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ the three rigged configurations $$\begin{array}{cc}& \\ & \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\hfill \\ & \\ & \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\hfill \\ & \\ & \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\hfill \end{array}\begin{array}{cc}& \\ & \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\hfill \end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\begin{array}{ccc}& & \\ & & \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\hfill \\ & & \\ & \multicolumn{2}{c}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}}\end{array}\begin{array}{cc}& \\ & \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\hfill \end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\text{and}}$}\begin{array}{ccc}& & \\ & & \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\hfill \\ & & \\ & \multicolumn{2}{c}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\end{array}\begin{array}{cc}& \\ & \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\hfill \end{array}$$ are restricted of level 2 with charges $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}`$, respectively. The riggings are given on the right of each part. Hence $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{K}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}`$. In contrast to this, the Kostka polynomial $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{K}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ is obtained by summing over both configurations in (5.8) with all possible riggings below the vacancy numbers. This amounts to $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{K}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{6}$}`$. In Section 7 we will use Theorem 5.7 to obtain explicit expressions for type $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}`$ branching functions. The results suggest that it is also useful to consider the following sets of rigged configurations with imposed minima on the set of riggings. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}`$ be a partition and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$, the sequence of single columns of height $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}`$. Set $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}`$ and $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}$$ for all $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{CST}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Then define $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RC}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ to be the set of all $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RC}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ such that there exists a $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{CST}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ such that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ for all $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. Note that the second condition is obsolete if $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$ occurs as a part in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ since by definition $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ for all $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$. Conjecture 8.3 asserts that the set $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RC}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ corresponds to the set of all level-$`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$ restricted Littlewood–Richardson tableaux with a fixed subtableaux of shape $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}`$. ## 6. Fermionic expression of level-restricted generalized Kostka polynomials ### 6.1. Fermionic expression Similarly to the Kostka polynomial case, one can rewrite the expression of the level-restricted generalized Kostka polynomials of Theorem 5.7 in fermionic form. ###### Lemma 6.1. For all $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{C}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{CST}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$, we have $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$. ###### Proof. Since $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$ it follows from \[26, (11.2)\] that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$. Since $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}`$ is over the alphabet $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}`$ this implies for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$ $$\begin{array}{cc}\hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\hfill \\ \hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}\hfill \end{array}$$ Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{SCST}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ be the set of all nonempty subsets of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{CST}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Furthermore set $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{min}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{SCST}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Then by inclusion-exclusion the set of allowed rigging for a given configuration $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{C}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ is given by $$\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{SCST}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ Since the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}`$-binomial $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left[}$}\genfrac{}{}{0.0pt}{}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right]}$}`$ is the generating function of partitions with at most $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}`$ parts each not exceeding $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}`$ and since $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ by Lemma 6.1 the level-$`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$ restricted generalized Kostka polynomials has the following fermionic form. ###### Theorem 6.2. $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{K}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{SCST}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{C}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left[}$}\genfrac{}{}{0.0pt}{}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right]}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ In Section 7 we will derive new expressions for branching functions of type $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}`$ as limits of the level-restricted generalized Kostka polynomials. To this end we need to reformulate the fermionic formula of Theorem 6.2 in terms of a so-called $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒎}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒏}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$-system. Set $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\eta }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ which is the number of rectangles in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$ of shape $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Then $$\begin{array}{cc}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\hfill \\ \hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\hfill \\ & \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\eta }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{min}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{min}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{min}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\hfill \\ & \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\hfill \\ & \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\hfill \\ \hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}\hfill \end{array}$$ At this stage it is convenient to introduce vector notation. For a matrix $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ with indices $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ define $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒗}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒆}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒆}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ where $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒆}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒆}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}`$ are the canonical basis vectors of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$, respectively. Define $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}$$ which in vector notation reads (6.1) $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒖}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{I}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒇}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒆}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒆}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ where $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}`$ is the Cartan matrix of type $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{I}$}`$ is the identity matrix. Since $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ by Lemma 6.1 it follows that (6.2) $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{I}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒎}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{I}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒏}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝑳}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒖}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ In terms of the new variables the condition (5.1) on $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}`$ becomes (6.3) $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒆}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒆}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{I}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒏}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{min}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ where we used $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{min}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$ if $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}`$ is $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\times }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$-dimensional and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}`$. ###### Lemma 6.3. In terms of the above $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝐦}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝐧}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$-system (6.4) $$\begin{array}{c}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒎}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒎}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒎}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{I}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒖}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\hfill \\ \hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒖}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒖}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\end{array}$$ where $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}$$ and $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{min}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$. ###### Proof. By definition $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cc}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}`$. Note that $$\begin{array}{cc}\hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\hfill \\ & \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\hfill \\ & \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{c}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{c}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}\hfill \end{array}$$ Hence eliminating $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cc}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ in favor of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}`$ yields $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ On the other hand, using $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$, $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒏}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{I}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{I}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝑷}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}$$ so that (6.5) $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒏}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{I}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{I}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝑷}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝑳}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ Eliminating $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒏}$}`$ in favor of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒎}$}`$ using (6.2) and substituting $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝑷}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒎}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒇}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ yields $$\begin{array}{c}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒏}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{I}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{I}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝑷}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝑳}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒎}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒎}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝑳}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒇}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{I}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝑳}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒖}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}\hfill \\ \hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝑳}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒖}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{I}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝑳}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒇}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}\end{array}$$ Similarly, replacing $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒏}$}`$ by $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒎}$}`$ in (6.3) we obtain (6.6) $$\begin{array}{c}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒆}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒆}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{I}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒎}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒖}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\hfill \\ \hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}\end{array}$$ Inserting these equations into (6.5), trading $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒇}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒖}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ by (6.1) and using $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{I}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝑳}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝑳}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒆}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒆}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}$$ results in the claim of the lemma. ∎ As a corollary of Lemma 6.3 and Theorem 6.2 we obtain the following expression for the level-restricted generalized Kostka polynomial (6.7) $$\begin{array}{c}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{K}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{SCST}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}^{\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒖}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒖}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\hfill \\ \hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\times }$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒎}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}^{\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒎}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒎}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒎}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{I}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒖}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left[}$}\genfrac{}{}{0.0pt}{}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒎}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒏}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒎}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right]}$}\end{array}$$ where $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒏}$}`$ is determined by (6.2), the sum over $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒎}$}`$ is such that $$\begin{array}{c}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒆}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒆}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{I}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒎}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒖}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\hfill \\ \hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\end{array}$$ for all $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left[}$}\genfrac{}{}{0.0pt}{}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒎}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒏}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒎}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right]}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left[}$}\genfrac{}{}{0.0pt}{}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right]}$}`$. Now consider the second case of Example 5.6, namely $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$. Then $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{SCST}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ only contains the element $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}`$ where $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}`$ is the Yamanouchi tableau of shape $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒖}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒆}$}_\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒆}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}`$. In the vacuum case, that is, when $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$, the set $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{SCST}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ only contains $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒖}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒇}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$. In this case (6.7) simplifies to $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{K}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒎}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}^{\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒎}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒎}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left[}$}\genfrac{}{}{0.0pt}{}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒎}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒏}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒎}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right]}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ When $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$ is a sequence of single boxes this proves \[8, Theorem 1\]<sup>1</sup><sup>1</sup>1We believe that the proof given in is incomplete.. When $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$ is a sequence of single rows or single columns this settles \[12, Conjecture 4.7\]. ### 6.2. Polynomial Rogers–Ramanujan-type identities Let $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}}`$ be the Weyl group of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}`$ be the root lattice, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}`$ the half-sum of the positive roots, and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ the standard symmetric bilinear form. Recall the energy function (3.9). It was shown in that (6.8) $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{K}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\underset{\begin{array}{c}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒫}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{wt}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\end{array}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ Equating (6.7) and (6.8) gives rise to polynomial Rogers–Ramanujan-type identities. For the vacuum case, that is, when the partition $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}`$ is rectangular with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ rows, this proves \[33, Eq. (9.2)\]<sup>2</sup><sup>2</sup>2The definition of level-restricted path as given in \[33, p. 394\] only works when $`\colorbox[rgb]{1,1,1}{$R$}`$ (or $`\colorbox[rgb]{1,1,1}{$\mu $}`$ therein) consists of single rows; otherwise the description of Section 3.7 should be used.. ## 7. New expressions for type $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}`$ branching functions The coset branching functions $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}`$ labeled by the three weights $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}`$ have a natural finitization in terms of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$-restricted crystals. For certain triples of weights these can be reformulated in terms of level-restricted paths, which in turn yield an expression of the type $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}`$ branching functions as a limit of the level-restricted generalized Kostka polynomials. Together with the results of the last section this implies new fermionic expressions for type $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}`$ branching functions at certain triples of weights. ### 7.1. Branching function in terms of paths Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cl}}$}}`$ be dominant integral weights of levels $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$, and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}`$ respectively, where $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}`$. The branching function $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ is the formal power series defined by $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{z}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{af}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{af}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{af}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}}$$ where $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{af}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{af}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{af}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}}`$ is the multiplicity of the irreducible integrable highest weight $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$-module $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝕍}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{af}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ in the tensor product $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝕍}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{af}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝕍}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{af}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. The desired multiplicity is equal to the number of $`\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$-highest weight vectors of weight $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{af}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}`$ in the tensor product $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝔹}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{af}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝔹}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{af}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$, that is, the number of elements $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝔹}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{af}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝔹}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{af}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ such that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{wt}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{af}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ for all $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{I}$}`$. By (3.5), $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ is $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}`$-restricted, and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{wt}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{af}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}`$. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ be a perfect crystal of level $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$. Using the isomorphism $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝔹}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$ and $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ be the ground state path. Suppose $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}`$ is such that for all $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$. Write $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$. In type $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ the period of the ground state path $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}`$ always divides $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$. Choose $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}`$ to be a multiple of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$, so that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}`$ and $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. Then the above desired highest weight vectors have the form $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝔹}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{af}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝔹}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{af}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. But there is an embedding $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝔹}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{af}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝔹}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{af}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝔹}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{af}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ defined by $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}}`$. With this rephrasing of the conditions on $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$ and taking limits, we have (7.1) $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\underset{\begin{array}{c}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\end{array}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{lim}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{z}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{z}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}$$ where $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ is given by $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}`$ is the energy function on finite paths. Our goal is to express (7.1) in terms of level-restricted generalized Kostka polynomials. We find that this is possible for certain triples of weights. Using the results of Section 6 this provides explicit formulas for the branching functions. ### 7.2. Reduction to level-restricted paths The first step in the transformation of (7.1) is to replace the condition of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$-restrictedness by level $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$ restrictedness. This is achieved at the cost of appending a fixed inhomogeneous path. Consider any tensor product $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}`$ of perfect crystals each of which has level at most $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}`$ (the level of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}`$), such that there is an element $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. We indicate how such a $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}`$ can be constructed explicitly. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}`$ be the partition with strictly less than $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ rows with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ columns of length $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}`$ be the Yamanouchi tableau of shape $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}`$. Then any factorization (in the plactic monoid) of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}`$ into a sequence of rectangular tableaux, yields such a $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}`$. ###### Example 7.1. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{6}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}`$. Then $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ (its transpose is $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ and $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\begin{array}{cccc}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}& & \\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}& & & \end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ One way is to factorize into single columns: $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ where each $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$ is an $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ highest weight vector, namely, the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$-th column of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}`$. Another way is to factorize into the minimum number of rectangles by slicing $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}`$ vertically. This yields $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$; again the factors of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ are the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ highest weight vectors, namely, $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\begin{array}{cc}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\begin{array}{c}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\begin{array}{c}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ Consider also a tensor product $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ of perfect crystals such that there is an element $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Then $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Instead of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$, we work with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}`$ where $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}`$ is restricted of level $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$. This trick doesn’t help unless one can recover the correct energy function directly from $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}`$. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}`$ be the first $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}`$ steps of the ground state path $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Define the normalized energy function on $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}`$ by $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. A priori it depends on $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$, and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$. The energy function occurring in the branching function is $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. ###### Lemma 7.2. $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$. ###### Proof. It suffices to show that the function $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ given by $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ is constant. Using the definition (3.9) and the fact that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$ is homogeneous of length $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}`$, we have $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ Similarly $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Therefore $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Thus it suffices to show that the function $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ given by $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ is a constant function. Suppose first that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ for some $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. By the construction of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ and $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$, since $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$. Then $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ by (3.4). Passing from $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ to $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ repeatedly, the values of the energy functions are constant, so it may be assumed that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ is a $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ highest weight vector; in particular, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ for all $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. Next suppose that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$. Now $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$. By (3.4) $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$. By (3.8) and the fact that the local isomorphism on $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ is the identity, we have $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. To show that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ we check the conditions of Lemma 3.2. By (3.1) $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{wt}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$. Also by (3.5), since $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$, we have $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}`$ be the image of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ under an arbitrary composition of local isomorphisms. Since $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ is an $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ highest weight vector, so is $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}`$. Now $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}`$ is the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$-highest weight vector in a perfect crystal of level at most $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$, so $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$. But $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ so that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$. By (3.4) $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}`$. So $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ by Lemma 3.2. By induction we may now assume that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$. But then $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$, or $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$. Since $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ is a perfect crystal of level $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ must be the unique element of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ such that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$. Thus the function $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ given by $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ is constant on $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ if it is constant on the singleton set $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}`$, which it obviously is. ∎ ### 7.3. Explicit ground state energy To go further, an explicit formula for the value $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ is required. This is achieved in (7.2). The derivation makes use of the following explicit construction of the local isomorphism. ###### Theorem 7.3. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}`$ be a perfect crystal of level $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cl}}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ a perfect crystal of level $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$, and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ (resp. $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$) be the unique element such that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}`$ (resp. $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$). Then under the local isomorphism $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$, we have $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}`$. The proof requires several technical lemmas and is given in the next section. ###### Example 7.4. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}`$. Here the set $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ consists of two elements, namely, $$\begin{array}{cc}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}\end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\text{and}}$}\begin{array}{cc}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}\end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$ be the second tableau. The theorem says that $$\begin{array}{cccc}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}\end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\begin{array}{cc}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}\end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\begin{array}{cc}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\begin{array}{cccc}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}\end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ ###### Proposition 7.5. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cl}}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}`$ a perfect crystal of level $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$, $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ the ground state path, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}`$ a finite path (say of length $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}`$ where $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}`$ is a multiple of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$) such that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ the tensor product of perfect crystals each of level at most $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$, and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ be the path of length $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}`$ such that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ where $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ is the ground state path. Then under the composition of local isomorphisms $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}`$ we have $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$. ###### Proof. Induct on the length of the path $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}`$. Suppose $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$ where $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$ is a perfect crystal. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{wt}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. By the definitions $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. By induction the first $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}`$ steps $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}`$ of the ground state path of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ satisfy $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ under the composition of local isomorphisms $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}`$. Tensoring on the left with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$, it remains to show that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}`$ under the composition of local isomorphisms $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}`$. Now $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ are the unique elements such that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$. Applying Theorem 7.3 we obtain $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}`$. Now $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ so that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. This implies that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Now by definition $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Applying Theorem 7.3 we obtain $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}`$. Continuing in this manner it follows that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. Composing these local isomorphisms it follows that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}`$. But $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}`$ is the identity since the order of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}`$ divides $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ which divides $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}`$. Therefore $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}`$ under the composition of local isomorphisms and we are done. ∎ In the notation in the previous section, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ where $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ is the first $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}`$ steps of the ground state path of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Write $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}`$. Then using the generalized cocyclage one may calculate explicitly the generalized charge of the LR tableau corresponding to the level $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ restricted (and hence classically restricted) path $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}`$ denote the total number of cells in the tableaux comprising $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$. Then (7.2) $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\genfrac{}{}{0pt}{}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ ###### Example 7.6. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. Then $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ is the path $$\begin{array}{ccc}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}\end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\begin{array}{ccc}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\begin{array}{ccc}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}\end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\begin{array}{ccc}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\begin{array}{ccc}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ The element $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ can be taken to be the tensor product $$\begin{array}{c}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\begin{array}{c}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{8}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{8}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{8}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{7}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{6}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Then the tableau $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{LR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ (resp. $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}`$) that records the path $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ (resp. $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$) is given by $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\begin{array}{cccccccc}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{11}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{15}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{7}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{7}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{7}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{12}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{16}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{8}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{8}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{8}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{13}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{17}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{9}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{9}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{9}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{14}$}& \\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{6}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{6}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{6}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}& & \end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\begin{array}{cc}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{6}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{7}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}& \end{array}$$ with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ and subalphabets $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{6}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{7}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{8}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{9}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{11}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{12}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{13}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{14}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{15}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{16}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{17}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}`$. The generalized charge $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ is equal to the energy $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ \[37, Theorem 23\]. Here the widest rectangle in the path is of width $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$. For any tableau $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{LR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ for some partition $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}`$, define $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ where $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}`$ is the Schensted $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}`$ tableau, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$ is the automorphism of conjugation that reverses each of the subalphabets, and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}`$ are the west and east subtableaux obtained by slicing $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}`$ between the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$-th and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$-th columns. It can be shown that there is a composition of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}`$ generalized $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$-cocyclages leading from $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}`$ to $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ where $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}`$ denotes the number of cells in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}`$. It follows from the ideas in \[35, Section 3\] and the intrinsic characterization of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$ in \[35, Theorem 21\] that (7.3) $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ For the above tableau $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}`$ we have $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\begin{array}{ccc}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{6}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{6}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{6}$}\end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\begin{array}{ccc}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}\end{array}$$ and $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\begin{array}{ccccc}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{11}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{15}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{7}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{7}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{7}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{12}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{16}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{8}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{8}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{8}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{13}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{17}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{9}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{9}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{9}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{14}$}& \\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}& & \end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\begin{array}{ccccc}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{6}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{6}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{6}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{11}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{15}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{7}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{7}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{7}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{12}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{16}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{8}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{8}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{8}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{13}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{17}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{9}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{9}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{9}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{14}$}& \\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}& & \end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ Then $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\begin{array}{ccccc}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{11}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{15}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{12}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{16}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{13}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{17}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{14}$}& \\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}& & \\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{6}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{6}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{6}$}& & \\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{7}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{7}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{7}$}& & \\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{8}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{8}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{8}$}& & \\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{9}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{9}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{9}$}& & \\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}& & \end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\text{and}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\begin{array}{ccc}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{6}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{6}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{6}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{7}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{7}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{7}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{8}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{8}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{8}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{9}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{9}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{9}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{11}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{15}$}& \\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{12}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{16}$}& \\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{13}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{17}$}& \\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{14}$}& & \end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ We have $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ by \[35, Theorem 21\] and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}`$, and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}`$ by (7.3). This implies $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}`$. ### 7.4. Proof of Theorem 7.3 The proof of Theorem 7.3 requires several lemmas. Words of length $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}`$ in the alphabet $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}`$ are identified with the elements of the crystal basis of the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}`$-fold tensor product $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}`$. ###### Lemma 7.7. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}`$ be words such that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}`$ is an $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ highest weight vector. Then $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}`$ is an $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ highest weight vector and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ for all $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. ###### Proof. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}`$ be an $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ highest weight vector and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. By (3.5) $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{max}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ Since both summands on the right hand side are nonnegative and sum to zero they must both be zero. ∎ ###### Lemma 7.8. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}`$ be a word in the alphabet $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}`$ and $`\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}}`$ a word obtained by removing a letter $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$ of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}`$. Then 1. $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ with equality only if $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. 2. $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ with equality only if $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$. ###### Proof. Write $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}`$ and $`\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}`$. By (3.5) (7.4) $$\begin{array}{cc}\hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{max}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}\hfill \\ & \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\begin{array}{cc}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{max}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}\hfill & \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\text{if }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\hfill \\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\hfill & \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\text{if }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\text{.}}$}\hfill \end{array}\hfill \end{array}$$ In particular $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. Applying (3.5) to both $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ and subtracting, we obtain $$\begin{array}{cc}\hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{max}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{max}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}\hfill \\ & \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{max}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{max}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}\hfill \\ & \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}\hfill \end{array}$$ Moreover if $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ then all of the inequalities are equalities. In particular it must be the case that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$, which by (7.4) implies that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$, proving the first assertion. On the other hand, (7.4) also implies $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Subtracting $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ from $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ and computing as before, the second part follows. ∎ Say that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}`$ is an almost highest weight vector with defect $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$ if there is an index $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ such that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$, and also $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ if $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. ###### Lemma 7.9. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}`$ be an almost highest weight vector with defect $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. Then $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ is either an $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ highest weight vector or an almost highest weight vector of defect $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. ###### Proof. For $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}`$, the restriction of the words $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ to the alphabet $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}`$ are identical, so that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ by the definition of an almost highest weight vector. Also $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ implies that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$. Again by the definition of an almost highest weight vector, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$. If $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ we have shown that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ is an $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ highest weight vector. So it may be assumed that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. It is enough to show that one of the two following possibilities occurs. 1. $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$. 2. $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$. Recall that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ is obtained from $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}`$ by changing an $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ into an $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$. Write $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}`$ such that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}`$. In this notation we have $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$. By Lemma 7.8 point 1 with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}`$ replaced by $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}`$ and using that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}`$ is an almost highest weight vector of defect $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$, we have $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. It is now enough to assume that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ and to show that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$. By (3.5) $$\begin{array}{cc}\hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\hfill \\ & \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{max}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}\hfill \end{array}$$ In particular $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$. Hence $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}`$. Similar computations starting with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ and which use the fact that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$, yield $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$. We have $$\begin{array}{cc}\hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\hfill \\ & \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{max}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}\hfill \\ & \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{max}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}\hfill \end{array}$$ But $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ and in passing from $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}`$ to $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ an $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$ is changed into an $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. By Lemma 7.8 point 2 applied to the restriction of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}`$ to the alphabet $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}`$, we have $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. It follows that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$, and that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ is an almost highest weight vector of defect $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. ∎ ###### Lemma 7.10. Suppose $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}`$ is an $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ highest weight vector and $`\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}}`$ is a word obtained by removing a letter (say $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$) from $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}`$. Then there is an index $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}`$ such that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ is an $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ highest weight vector. ###### Proof. By Lemma 7.9 it suffices to show that $`\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}}`$ is either an $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ highest weight vector or an almost highest weight vector of defect $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$. First it is shown that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$. For $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}`$, the restrictions of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}`$ and $`\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}}`$ to the alphabet $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}`$ are the same, so that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$. For $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$, by Lemma 7.8 point 1 and the assumption that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}`$ is an $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ highest weight vector, it follows that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. But equality cannot hold since the removed letter is $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$ as opposed to $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. Thus $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$. Next we observe that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ by Lemma 7.8 point 1 and the fact that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}`$ is an $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ highest weight vector. If $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ then $`\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}}`$ is an $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ highest weight vector. So it may be assumed that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. It suffices to show that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$. Write $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}`$ and $`\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}`$. Now $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ for all $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ by Lemma 7.7 since $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}`$ is an $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ highest weight vector. In particular $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ so that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}`$. We have $$\begin{array}{cc}\hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\hfill \\ & \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{max}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}\hfill \\ & \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{max}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}\hfill \end{array}$$ since $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ by Lemma 7.7. It is enough to show that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. But $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ The first inequality holds by an application of Lemma 7.8 point 2 since the restrictions of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ to the alphabet $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}`$ differ by inserting a letter $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$. The last inequality holds by Lemma 7.7 since $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}`$ is an $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ highest weight vector. ∎ ###### Lemma 7.11. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}`$ be a perfect crystal of level $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cl}}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ a finite (possibly empty) tensor product of perfect crystals of level at most $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ such that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}`$ such that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ and set $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$. Then there is an index $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}`$ such that (7.5) $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}$$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ where the subscripts are taken modulo $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$. Moreover if $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$ then $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}`$. ###### Proof. Since the Dynkin diagram $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ has an automorphism given by rotation, it may be assumed that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}`$ be the partition of length less than $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$, given by $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$. Since $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ it follows that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}`$ has a column of size $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$ be the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$-highest weight vector in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}`$. Write $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ and $`\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. Observe that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}}`$ is an affine highest weight vector in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝔹}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ and has weight $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}`$ so its connected component is isomorphic to $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝔹}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. A similar statement holds for $`\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝔹}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. In particular, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}`$ is an $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ highest weight vector. The map $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{word}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{word}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{word}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ gives an embedding of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$-crystals into a tensor product of crystals $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$. By Lemma 7.10, there exists an index $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ such that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{word}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{word}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{word}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ is an $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ highest weight vector. Since $`\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}}`$ is an $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ highest weight vector it follows that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{word}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{word}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{word}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{word}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{word}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{word}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$ be the position of the letter in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{word}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{word}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ that changes from a $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ to $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$ upon the application of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$, for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. It follows from the proof of Lemma 7.9 that (7.6) $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}`$ be the maximal index such that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}`$ is located in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{word}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Write $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. It follows that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ and that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}}`$ is an $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ highest weight vector. It remains to show that (7.7) $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}$$ and that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}`$ with equality if $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$. Consider the corresponding positions in the tableau $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$. Since $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{word}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ is an $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$-crystal morphism, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{word}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{word}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ be the position in the tableau $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$ corresponding to the position $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{word}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$, and analogously define $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$, and so on. Since the rows of all tableaux (and in particular $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$, etc.) are weakly increasing and (7.6) holds, it follows that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}`$. But $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$ has $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}`$ rows, so $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}`$. The next goal is to prove (7.7). Suppose first that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. In this case the letters $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ are undisturbed in passing from $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ to $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Using this and the Dynkin diagram rotation it follows that (7.8) $$\begin{array}{cc}\hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\hfill \\ & \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{max}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}\hfill \\ & \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{max}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}\hfill \end{array}$$ But $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ by the fact that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ and Lemma 7.8 point 2 applied after rotation of the Dynkin diagram. By (7.8) the desired result (7.7) follows. Otherwise assume $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. Here $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ since $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ with the inequality holding by the perfectness of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$. By (7.6) and the fact that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$ is a tableau, it must be the case that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ acting on $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$ changes a $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$ in the first row of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$ into a $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$ acting on $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ changes a $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}`$ in the second row of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ into a $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$, etc. Since $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$ is a tableau with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ rows with entries between $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$, there are integers $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ such that the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$-th row of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$ consists of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$ copies of the letter $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$ copies of the letter $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. For tableaux $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$ of this very special form, the explicit formula for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ in \[37, (3.11)\] yields $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ where $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ is the number of occurrences of the letter $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$. Since $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ also has the same form (with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$ replaced by $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$) and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$, it follows that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. We have $$\begin{array}{cc}\hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\hfill \\ & \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{max}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}\hfill \\ & \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{max}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\hfill \end{array}$$ since $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Finally, assuming $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$, it must be shown that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}`$. Since the level of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ is the same as that of the weights $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$, it follows from the perfectness of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ that both $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ are uniquely defined by the property that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{z}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$. By the explicit construction of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$ in Example 3.3 $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{wt}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{z}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{z}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}$$ with indices taken modulo $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$. Subtracting the analogous formula for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{wt}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{wt}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{wt}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$. Using (3.1) it follows that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}`$. ∎ ###### Proof of Theorem 7.3. First observe that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ by (3.1), $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$, and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}`$. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ be such that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{z}$}`$ under the local isomorphism. Then $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ which means that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{z}$}`$ is $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$-restricted. Hence $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. The former together with the perfectness of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ implies that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{z}$}`$. From the latter it follows that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. However the set $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ might have multiplicities so it is not obvious why $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ or equivalently $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. The proof proceeds by an induction that changes the weight $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ to a weight $`\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}`$ that is “closer to” $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$. Suppose first that there is a root direction $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ such that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ and $`\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$. By Lemma 7.11 applied for the weight $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$, simple root $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$, and element $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$, there is an $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ such that $`\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ where $`\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}`$. Applying Lemma 7.11 with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}}`$, and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$, it follows that $`\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. The above computations imply $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. We have $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ since $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}`$ under the local isomorphism. It must be seen which of these raising operators act on the tensor factor in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ and which act in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$. By Lemma 7.11 applied with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$, and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$, it follows that $`\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ and that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}}`$. Since $`\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}_{\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}}`$ is an $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ highest weight vector, the rest of the raising operators $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}}`$ must act on the first tensor factor. Let $`\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Then $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}}`$. But the local isomorphism is a crystal morphism so it sends $`\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}}`$. By induction $`\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\widehat{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. By (3.6) it follows that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Otherwise there is no index $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ such that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$. This means $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$. But the sets $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ are singletons whose lone elements are given by the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ highest weight vectors in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ respectively. Since $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ is $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ multiplicity-free it follows that the sets $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ are singletons. In this case it follows directly that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ since both $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ are elements of the set $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. ∎ ### 7.5. Branching function by restricted generalized Kostka polynomials The appropriate map from LR tableaux to rigged configurations, sends the generalized charge of the LR tableau to the charge of the rigged configuration. Unfortunately in general it is not clear what happens when one uses the statistic coming from the energy function $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ but using the path $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}`$. It is only known that the statistic $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ on the path $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}`$, is well-behaved. So to continue the computation we require that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$. This is achieved when $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$. So let us assume this. The other problem is that we do not consider all paths in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$, but only those of the form $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ where $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ is a fixed path. Passing to LR tableaux, this is equivalent to imposing an additional condition that the subtableaux corresponding to the first several rectangles, must be in fixed positions. Conjecture 8.3 asserts that the corresponding sets of rigged configurations are well-behaved. The special case that requires no extra work, is when $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ consists of a single perfect crystal. This is achievable when $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ has the form $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$; in this case $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ is the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$-highest weight element of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}}`$. This is the same as requiring that the first subtableau of the LR tableau be fixed. But this is always the case. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ consist of the single rectangle $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ followed by $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ copies of the rectangle $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}_{}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}{}_{}{}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ where $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}`$. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ be the partition of the same size as the total size of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$, such that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ projects to $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$. Then the set of paths $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ is equal to $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒫}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}`$. This is summarized by (7.9) $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{lim}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\genfrac{}{}{0pt}{}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{K}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ where $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}`$ is arbitrary, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$, and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$. Inserting expression (6.7) for the generalized Kostka polynomial in (7.9) and taking the limit yields the following fermionic expression for the branching function (7.10) $$\begin{array}{c}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}^{\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{SCST}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}^{\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒖}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒖}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\hfill \\ \hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\times }$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒎}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}^{\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒎}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒎}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒎}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{I}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒖}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\underset{\begin{array}{c}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\end{array}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left[}$}\genfrac{}{}{0.0pt}{}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right]}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\end{array}$$ where $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}`$ is any partition which projects to $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒖}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ as defined in (6.1). The sum over $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒎}$}`$ runs over all $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒎}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒆}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒆}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}`$ such that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ and $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒆}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒆}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{I}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒎}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒖}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}$$ for all $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. The variables $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ are given by $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒆}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒆}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒎}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{I}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒖}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒆}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒆}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right\}}$}$$ for all $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$. ## 8. Proof of Theorem 5.7 To prove Theorem 5.7 it clearly suffices to show that there is a bijection $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RLR}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RC}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ that is charge-preserving, that is, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ for all $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RLR}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Here we identify $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{LR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RLR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ via the standardization bijection $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{std}}$}`$. Also define $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{CLR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ by $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$ where $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RLR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$. It will be shown that one of the standard bijections $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RLR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RC}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ is charge-preserving, and that it restricts to a bijection $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RLR}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RC}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. With this in mind let us review the bijections from LR tableaux to rigged configurations. ### 8.1. Bijections from LR tableaux to rigged configurations A bijection $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{CLR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RC}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ was defined recursively in \[25, Definition-Proposition 4.1\]. It is one of four natural bijections from LR tableaux to rigged configurations. 1. Column index quantum: $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{CLR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RC}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. 2. Column index coquantum: $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{CLR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RC}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$, defined by $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$. 3. Row index quantum: $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RLR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RC}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$, defined by $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tr}}$}`$, and 4. Row index coquantum: $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RLR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RC}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$, defined by $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$. Of these four, the one that is compatible with level-restriction is $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}}`$. First we show that it is charge-preserving. This fact is a corollary of the difficult result \[25, Theorem 9.1\]. ###### Proposition 8.1. $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ for all $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RLR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. ###### Proof. Consider the following diagram, which commutes by the definitions and \[25, Theorem 7.1\] In particular $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tr}}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RC}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RLR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Then, using $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tr}}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RC}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tr}}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RC}}$}}`$, $$\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tr}}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RC}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tr}}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RC}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Then $$\begin{array}{cc}\hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cc}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\hfill \\ & \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cc}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tr}}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RC}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cc}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\hfill \end{array}$$ by Lemma 5.4, (5.6) and \[25, Theorem 9.1\] to pass from $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cc}}$}`$ to $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$. ∎ In light of Proposition 8.1, to prove Theorem 5.7 it suffices to establish the following result. ###### Theorem 8.2. The bijection $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RLR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RC}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ restricts to a well-defined bijection $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RLR}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RC}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Computer data suggests that the bijection $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$ is not only well-behaved with respect to level-restriction, but also with respect to fixing certain subtableaux. It was argued in Section 7.5 that the branching functions can be expressed in terms of generating functions of tableaux with certain fixed subtableaux. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}`$ be partitions, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}`$ the unique tableau in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RLR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Define $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RLR}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ to be the set of tableaux $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RLR}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ such that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}`$ restricted to shape $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}`$ equals $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}`$. Recall the set of rigged configurations $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RC}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ defined in Section 5.3. ###### Conjecture 8.3. The bijection $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RLR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RC}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ restricts to a well-defined bijection $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RLR}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RC}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. ### 8.2. Reduction to single rows In this section it is shown that to prove Theorem 8.2 it suffices to consider the case where $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$ consists of single rows. Recall the nontrivial embedding $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{LR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{LR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{r}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. We identify $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{LR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RLR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ via $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{std}}$}`$, and therefore have an embedding $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RLR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RLR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{r}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Define a map $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RC}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RC}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{r}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ as follows. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RC}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. For each rectangle of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$ having $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}`$ rows and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}`$ columns, add $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$ strings $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ of length $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}`$ and label zero to the rigged partition $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. The resulting rigged configuration is $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. ###### Proposition 8.4. The following diagram commutes: $$\begin{array}{ccc}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RLR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}& \stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RLR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{r}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\\ \overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}& & \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{r}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}}& & \\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RC}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}& \underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RC}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{r}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}\end{array}$$ It must be shown that similar diagrams commute in which $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$ is replaced by either $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}`$ or $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}`$, the maps that occur in the definition of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RC}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RC}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ be defined by adding a string $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ to each of the first $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\eta }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ rigged partitions in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RC}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. ###### Lemma 8.5. $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}`$ is well-defined and the following diagram commutes: $$\begin{array}{ccc}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RLR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}& \stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RLR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\\ \overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}& & \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}}& & \\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RC}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}& \underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RC}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}\end{array}$$ ###### Proof. Consider the following diagram. Let us view this diagram as a prism in which the large rectangular face is the front, and the other faces with four sides are the top and bottom, and the faces with three sides are the left and right. We want to show that the front face commutes. For this it suffices to show that all other faces commute. The left and right faces commute by the definition of $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}}`$. Let us define $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{CLR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{CLR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_{}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}}{}_{}{}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ so that the top face commutes. It suffices to show the bottom face commutes. Observe that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ is the embedding for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{CLR}}$}`$ that splits off the first column of the first rectangle in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}`$. But then the bottom face commutes by \[25, Lemma 5.4\] applied to $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}`$ in place of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$. ∎ ###### Lemma 8.6. The following diagram commutes: $$\begin{array}{ccc}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RLR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}& \stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RLR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\\ \overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}& & \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}}& & \\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RC}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RC}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}\end{array}$$ ###### Proof. We use the same kind of diagram as in the previous lemma. Of course $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. We argue as in the previous lemma. The left and right faces commute by the definition of $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}}`$, the top face commutes by \[36, Proposition 32\], and the bottom face commutes by \[25, Lemma 8.5\]. ∎ ###### Proof of Proposition 8.4. Consider the diagram The left and right faces commute by the definition of $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}}`$. Let us define $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{I}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ so that the top face commutes. It suffices to show the bottom face commutes. By the previous two lemmas, the bottom face commutes if $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$ is given by the composition of maps of the form $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}`$ and the identity map, corresponding to the way that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$ was computed. But it is easy to see that the effect of this composition of maps is precisely $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$. ∎ By the definition of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$ and Definition 5.5 of the level-restriction for rigged configurations, we have (8.1) $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RC}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RC}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RC}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{r}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ We now show that Theorem 8.2 follows from the special case when $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$ consists of single rows. The proof is a diagram chase using the the commutative diagram in Proposition 8.4. Since $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{r}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ consists of single rows, it is assumed that $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{r}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{LR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{r}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RC}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{r}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ restricts to a bijection $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{LR}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{r}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RC}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{r}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. In particular $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{r}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{LR}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{r}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RC}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{r}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Since $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{LR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RC}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ is a bijection, it is enough to show that $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{LR}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RC}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. For the inclusion $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{LR}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RC}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$, suppose that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{LR}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. By (4.7) $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{LR}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{r}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. By assumption, $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{r}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RC}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{r}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. But $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{r}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$ by Proposition 8.4, so $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RC}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{r}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. By (8.1), $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RC}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. For the other inclusion, suppose $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RC}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{LR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ be the unique element such that $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}`$. Now $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{r}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. By (8.1) $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RC}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{r}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. By assumption $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{LR}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{r}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. By (4.7) $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{LR}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$, that is, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{LR}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. ### 8.3. Single row quantum number bijection We must prove Theorem 8.2 when $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$ consists of single rows. For the rest of the paper we shall assume this is the case. Then $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\eta }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ for all $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{LR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{CST}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{LR}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{CST}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$, and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}`$ consists of single columns. We also write $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}`$ for $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$ in this case. Again using $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{std}}$}`$ we identify $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{LR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RLR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$, and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{LR}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ with its image in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RLR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ under $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{std}}$}`$. Now \[25, Section 4.2\] gives a direct description of $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}`$ that is particularly simple when $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}`$ consists of single columns. This is easily translated to the following algorithm to compute the bijection $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RLR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RC}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. First, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{C}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ requires that $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}$$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. The vacancy numbers may be given by $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}$$ where $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}`$ and (since $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}`$ is not necessarily a partition) $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:=}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{min}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ Now let us describe the bijection $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RLR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RC}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Start with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RLR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Write $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ for the tableau obtained by removing the largest letter from $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}`$ (which occurs in row $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}`$, say) and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ for the shape of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Since $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RLR}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$, by induction $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{J}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ is defined. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$. For $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ down to $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$, select the longest singular string in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{J}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ of length $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ (possibly of zero length) such that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$. With the convention $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$, it can be shown that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ as well. Then $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ is obtained from $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{J}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ by lengthening each of the selected strings by one, and resetting their labels to make them singular with respect to the vacancy numbers in the definition of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RC}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$, and leaving all other strings unchanged. Denote the transformation $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{J}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ by $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. The inverse of $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$, denoted $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}}`$, is obtained as follows. Set $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}`$. Select inductively a singular string of length $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ smallest such that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$. If no such singular string exists set $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$. Then $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{J}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ is obtained from $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ by shortening all selected strings by one, making them singular again and leaving all other strings unchanged. ###### Remark 8.7. Up to the relabeling bijection $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{std}}$}`$ this is precisely the description of the bijection $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{CST}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{RC}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{;}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ that was given in terms of the map called $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pi }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ in . ###### Example 8.8. Take $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ and $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\begin{array}{ccc}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{6}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{8}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{9}$}& \\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{7}$}& & \end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\text{so that}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\begin{array}{ccc}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{6}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{8}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}& & \\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{7}$}& & \end{array}$$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}`$. The rigged configuration corresponding to $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ is $$\begin{array}{cccc}& & & \\ \hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}& & & \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\hfill \\ & & & \\ \hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}& & & \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\hfill \\ & & & \\ \hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}& \multicolumn{2}{c}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}}\end{array}\begin{array}{ccc}& & \\ \hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\hfill \\ & & \\ \hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}& & \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\hfill \end{array}\begin{array}{ccc}& & \\ \hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}& & \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\hfill \end{array}$$ where the labels are written to the right of each part and the vacancy numbers to the left. The selected strings under $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}`$ are indicated by $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$. Hence the rigged configuration corresponding to $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}`$ is $$\begin{array}{cccc}& & & \\ \hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}& & & \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\hfill \\ & & & \\ \hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}& & & \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\hfill \\ & & & \\ \hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}& & & \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\hfill \end{array}\begin{array}{cccc}& & & \\ \hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}& & & \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\hfill \\ & & & \\ \hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}& & \multicolumn{2}{c}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}}\end{array}\begin{array}{ccc}& & \\ \hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}& & \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\hfill \end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ ### 8.4. Proof of the single row case Now we come to the proof of Theorem 8.2 when $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$ is a sequence of single rows. More precisely we will prove the following theorem. ###### Theorem 8.9. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ be a partition of level $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ an array of positive integers not exceeding $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$. Then $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ is in the image of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{CST}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ under $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}`$ if and only if 1. $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$ for all $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$, and 2. there exists a column-strict tableau $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{CST}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ such that (8.2) $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}$$ for all $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$. ###### Remark 8.10. The first column of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{CST}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ has length $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$. Since $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}`$ is a column-strict tableau over the alphabet $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}`$ this requires that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$. ###### Remark 8.11. Since $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ the bounds in (8.2) can be rewritten as (8.3) $$\begin{array}{c}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\hfill \\ \hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}\end{array}$$ For the proof of Theorem 8.9 it will be useful to have the following graphical description of (8.3) in mind. Consider $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ strips of length $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$ and height $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ arranged on top of each other. Assign the label $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}`$ to the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}`$-th strip from the top. Within each strip assign a height label with height $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ at the bottom of the strip and height $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ at the top of the strip. Call the coordinate along the horizontal axis the position. Then draw a horizontal line from position $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}`$ to position $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ at height $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$ in the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}`$-th strip with a closed dot at position $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}`$ and an open dot at position $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ to indicate that the first position belongs to the line, whereas the second one does not. If $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ draw an open dot. If $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ does not exist draw a horizontal line from position $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}`$, indicated by a closed dot, to position $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$. If there is an open dot at position $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ of height $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$ in strip $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}`$, then there is also a dot at the same position and height in strip $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. Connect all such dots by a vertical line. This way one obtains $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ paths which all end at position $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$. The $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$-th path in strip 1 starts at position $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$ by Remark 8.10. Furthermore, since $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}`$ the paths do not intersect. The $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}`$-th strip contains $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ paths. The other $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}`$ paths already ended at position $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$ in previous strips. An example for a set of such nonintersecting paths is given in Figure 1. It corresponds to $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{8}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{6}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ and $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\begin{array}{cccc}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}& \\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}& & \\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}& & & \end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ The dashed lines separate the various strips. To read off the bound on $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ from the picture, draw a vertical line at position $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$. Suppose that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}`$ paths cross this line horizontally in strip $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}`$ (when the vertical line goes through a closed/open dot we consider this as crossing/not crossing). Then $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}`$ is the maximal possible rigging for strings of length $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$ in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$. For example, the vertical line at position 6 in Figure 1 crosses one line in strip 1, no line in strip 2 and 4, and two lines in strip 3, so that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{6}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{6}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{6}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{6}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{6}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{6}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$, and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{6}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{6}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Recall that the rigging of a singular string of length $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$ in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ equals the vacancy number $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Hence the above graphical description of the bounds shows that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ cannot contain singular strings of length $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$ in the intervals $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$, and $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$ if $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}`$, since in these intervals a vertical line at position $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$ would cross at least one path. Conversely, if $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$ is the length of a singular string in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ then it must be in the complements of these intervals, that is, (8.4) $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}`$ or $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$, or $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}`$ if $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}`$, or $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$ if $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}`$. Since $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ these intervals are pairwise disjoint, but some of these intervals can of course be empty. Graphically the conditions in (8.4) require that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$ lies between two paths. More precisely, the first case in (8.4) states that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$ lies to the left of the first path, the second condition requires that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$ lies between the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$-th and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$-th path, and the third case applies if there are more than $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ paths in the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}`$-th strip in which case $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$ lies between paths $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. The last condition applies if there are exactly $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ paths in strip $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}`$. None of these ends at $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$ in this strip and the condition implies that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$ lies to the right of the rightmost path. ###### Remark 8.12. We use the following conventions throughout the proof: $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$. Without further ado we present the gory details of the proof of Theorem 8.9. ### Proof of Theorem 8.9 We prove the theorem by induction on $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}`$. The theorem is true for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$ since then $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$. In this case $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}`$ is of level $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ and conditions 1 and 2 are trivially satisfied. ### Proof of the forward direction Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{CST}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ its image under the row-wise quantum number bijection. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ be the tableau obtained from $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}`$ by removing the rightmost largest entry. Set $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{shape}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{J}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ and denote by $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}`$ the row index of the cell $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}`$. Set $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{shape}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. The tableau $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ is of level $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$ since $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$ by the condition that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}`$ is of level $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$. By induction the theorem holds for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ so that $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$ and there exists a column-strict tableau $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}`$ of shape $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}`$ such that by (8.3) (8.5) $$\begin{array}{c}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\hfill \\ \hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\overline{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}\end{array}$$ for all $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$. Here $`\overline{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$, and $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ is the largest part of $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{J}}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ and zero if $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{J}}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ is empty. The aim is to show that conditions 1 and 2 of the theorem hold for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Denote by $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ the length of the selected singular string in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{J}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ under $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. By definition $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}`$. We claim that there exist indices $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}`$ such that (8.6) $`\overline{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\overline{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\text{for }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}`$ (8.7) $`\overline{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\overline{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ and (8.8) $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ where by definition $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}}`$. The proof proceeds by descending induction on $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}`$. We make frequent use of (8.4) applied to $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{J}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ where the first and third line are viewed as the cases $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ of the general interval appearing in the second line of (8.4). First assume $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. Note that $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ since $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$. Hence the existence of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ follows from (8.4) since the last case does not apply. In particular we have $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}`$. Now consider $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ and assume that $`\overline{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ for some $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$. Then by induction and the column-strictness of $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}`$ $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\overline{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}$$ which is a contradiction. Hence $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$. Since $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ is the length of a singular string and by induction $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ must be in the first or second set of the intervals in (8.4) with all quantities replaced by their barred counterparts which proves (8.6). Equation (8.7) follows since $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$ by Remark 8.10. Let us now prove condition 1 of the theorem. By construction $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ is obtained from $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{J}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ by increasing the length of the selected strings in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{J}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ by one for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}`$, making them singular again and leaving all other strings unchanged. For $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ this means that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{J}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ so that condition 1 of the theorem is satisfied by induction. Now assume $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. Since $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}`$ it follows from (8.6) with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\overline{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$. Hence $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$ which ensures condition 1 of the theorem for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$. It remains to prove that the second condition of the theorem holds. The vacancy numbers of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}`$ and $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}`$ are related as follows $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ By construction $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}`$. Hence (8.9) $$\begin{array}{c}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\hfill \\ \hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\end{array}$$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. In the remainder of the proof of the forward direction it will be shown that (8.2) holds for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ and that (8.9) implies (8.2) for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. We distinguish the cases $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$, and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$. ### Case $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. In this case $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$, $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$, $`\overline{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ and $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}`$ is a column-strict tableau over the alphabet $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}`$. Furthermore $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. By (8.9) we have for all $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$ (8.10) $$\begin{array}{c}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\hfill \\ \hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}\end{array}$$ Remark 8.10 requires that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$. Hence (8.11) $$\begin{array}{c}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\hfill \\ \hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}\end{array}$$ If $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ the term $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ vanishes. In this case set $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ which defines a column-strict tableau of shape $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}`$ over $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}`$. Then (8.10) implies (8.2). The case $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ is considerably harder to establish due to the extra term $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Our strategy is as follows. The term $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ can be absorbed by defining $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}`$ appropriately except in certain cases. In general this introduces extra terms for the bounds at $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$. These in turn can be absorbed by defining $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}}`$ appropriately (except in certain cases) and so on. If all $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ can be defined and all bounds for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ are written in the form of (8.2) we are done. In the exceptional cases (when (8.10) does not imply (8.2)) it can be shown that the corresponding tableau $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}`$ is not of level $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$ which contradicts the assumptions. Let us now plunge into the details. Define $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ and set $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. Since $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$, we have $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ equals the right-hand side of (8.10). Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ be the minimal index $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{[}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{]}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{[}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{]}$}`$ such that $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$, where $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}}`$ and $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$. If no such $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$ exists set $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. By definition $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ for $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}`$ so that we can sharpen the bounds in (8.10) for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ by adding $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Note that $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}`$. The case $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}`$ will be dealt with later. Suppose that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}`$. Then one finds using (8.11) (8.12) $$\begin{array}{c}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\hfill \\ \hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}\end{array}$$ Define $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{min}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}`$ recursively by descending $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$. From its definition it is clear that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ lies in the interval $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{[}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{]}$}`$. By descending induction on $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$ it also follows that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{[}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{]}$}`$ and that either $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ or $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. The latter case only occurs when $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. In addition there must exist an index $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$ such that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ if $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. This is because $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ is at its lower bound in the interval $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{[}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{]}$}`$ and this can happen in only two ways; either $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ which proves the assertion with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ or $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ in which case the assertion must be true by induction since the initial case is $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$. Note that it also follows by induction that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ if $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ is minimal. From its definition it follows that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}`$ and furthermore $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$ which are the conditions for column-strictness for the first two columns of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}`$. Hence (8.12) yields (8.2) for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. We proceed inductively on $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$. Assume that by induction $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{[}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{]}$}`$ is already defined for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}`$. In terms of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}`$ (8.10) reads (8.13) $$\begin{array}{c}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\hfill \\ \hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}\end{array}$$ For $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ define $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ to be the minimal index $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{[}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{]}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{[}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{]}$}`$ such that $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. If no such $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$ exists set $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. Note that $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$. The case $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ will be dealt with later. Now assume that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$, and define recursively $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{min}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}`$ on descending $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$. By definition $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{[}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{]}$}`$. As in the case $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ it follows by descending induction on $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$ that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{[}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{]}$}`$ and that either $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ or $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. By definition $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$. Let us now show that also $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ which would prove the column-strictness of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}`$. By definition $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}`$ for some $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$. Assume $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$. Then $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ which violates the column-strictness of $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}`$. Hence $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$. If $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ then $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}`$ as desired. If $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ then a problem can only occur if $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}`$. However in this case $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}`$ which is a contradiction. This proves $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$. By the same arguments as in the case $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ there must exist an index $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$ such that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ if $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. For minimal $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ it follows again by induction that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. Since by definition $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ for $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}`$ for some $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$, one can add $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}$$ to (8.13). If $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ then the sum of this term and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ does not exceed $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ and (8.2) is proven for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$. It remains to treat the case when there exists a $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ such that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\kappa }$}`$ be minimal with this property. We will show that in this case $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}`$ is not of level $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$ which contradicts the assumptions. We claim that there exist indices $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\kappa }$}`$ such that (8.14) $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}`$ (8.15) $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\kappa }$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$. The inequalities (8.14) and (8.15) hold for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\kappa }$}`$ with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\kappa }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\kappa }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ and some $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\kappa }$}`$ by the definition of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\kappa }$}`$. Now suppose that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\kappa }$}`$ and that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\kappa }$}`$ satisfying (8.14) and (8.15) have been defined by induction. Recall that either $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ or $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. First assume that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$. This implies in particular that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ by (8.15) and hence $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ by the definition of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$. Set $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ and choose $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}`$ such that (8.15) holds which must be possible by the definition of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$. Also $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ since otherwise $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ by the column-strictness of $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}`$ which yields a contradiction since then (8.14) and (8.15) cannot hold simultaneously. Next assume $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ be minimal such that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$; the existence of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}`$ was proved before. In addition it was shown that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. The existence of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}`$ follows again from the definition of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$. As before $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$. By definition $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Since $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ is given by the right-hand side of (8.13), it follows from (8.14), (8.15) and the fact that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{[}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{]}$}`$ that (8.16) $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\text{for }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\kappa }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ Define $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}`$ with corresponding rigged configurations $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$ be the row index of the cell $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$. Denote the length of the selected string in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ under $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}}`$ by $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$. We claim that (8.16) implies (8.17) $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\text{for }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\kappa }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\text{.}}$}$$ This is shown by induction on $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}`$. By construction $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$. In addition (8.18) $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ If $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ then (8.17) follows immediately since $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$. Hence assume that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$. Since $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$, the vacancy number at $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ is decreased by one with each application of $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}}`$ until $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ because of (8.18) at $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. By (8.16) it takes $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ applications of $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}}`$ until there is a singular string of length $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ in the first rigged partition. Since $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ by (8.15) and Remark 8.10 this means that the singular string occurs at $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ which proves (8.17) at $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. Now consider the cases $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\kappa }$}`$ and assume that (8.17) holds for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}`$. First assume that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$. In this case $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}`$ and by (8.15) $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ so that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ by (8.17) at $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. If $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ there is nothing to show since $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}`$. Hence assume that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$. Again by (8.16) and (8.18) it takes $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}`$ applications of $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}}`$ until there is a singular string of length $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ in the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}`$-th rigged partition. Since $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}`$ the singular string occurs at $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}`$ which proves (8.17). Next assume that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. Then $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}`$ so that by (8.15) $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}`$. Since $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ it takes at most $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}`$ applications of $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}}`$ before there is a singular string in the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$-th rigged partition of length not exceeding $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$. After that, by (8.16) and (8.18), it takes $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}`$ applications of $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}}`$ until there is a singular string of length $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ in the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}`$-th rigged partition. Hence altogether the existence of a singular string of length $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ is assured at $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}`$ which concludes the proof of (8.17). Recall that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\kappa }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\kappa }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. Therefore (8.17) implies that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\kappa }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ is finite for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\kappa }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. If $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ is finite this means that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}`$ so that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\kappa }$}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\kappa }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. Since at most $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\kappa }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ boxes can be removed from $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}^{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\kappa }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ in rows with index $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\kappa }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ it follows that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$. This implies $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ where $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{shape}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Hence $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ which contradicts the assumption that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}`$ is of level $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$. This concludes the proof of the case $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. ### Case $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$. In this case $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$, $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}`$ and $`\overline{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}`$. It is convenient to introduce $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$, (8.19) $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\begin{array}{cc}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{max}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}\hfill & \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\text{for }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\text{,}}$}\hfill \\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\hfill & \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\text{for }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\text{,}}$}\hfill \end{array}$$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$. Note that for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}`$ either $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ or $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ and that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$. Define $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$, (8.20) $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\begin{array}{cc}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\hfill & \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\text{for }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\text{ and }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\text{,}}$}\hfill \\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\hfill & \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\text{for }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\text{,}}$}\hfill \\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{max}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}\hfill & \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\text{for }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\text{,}}$}\hfill \end{array}$$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$. By (8.6) we have $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}`$ so that (8.21) $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\text{for }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\text{.}}$}$$ It needs to be shown that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}`$ indeed defines a column-strict tableau over the alphabet $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}`$. Since $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$ for all $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}`$ the condition $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}`$ might only be violated if $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}`$. By (8.6) the latter condition requires $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ so that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ by (8.8). Since $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}`$ the first condition in (8.19) applies for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}`$. However, since $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ this implies that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ which contradicts the requirement $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$. This shows that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}`$. Next we check that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}`$ is column-strict. The condition $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}`$ only needs to be checked for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ since in all other cases it automatically follows from the column-strictness of $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}`$. First assume $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$. Then $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}`$ by (8.6). Furthermore $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}`$ by (8.6) and (8.20). Next assume $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. Then for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}`$ by (8.6) and (8.21). For $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ the column-strictness is trivial. Furthermore $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{max}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}`$ and $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{max}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}`$ so that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$. And finally $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{max}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}`$. The conditions $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ only need to be verified for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$, for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}`$, and for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}`$. First assume $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. Then $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{max}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$. Now assume $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}`$. For $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ we have $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ by (8.6). For $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ we have $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$. For $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ one obtains $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{max}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$. Next assume $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$. Then the second case of (8.19) applies so that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$. By (8.6) also $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}`$. This implies $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{max}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$. And finally for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ we have $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{max}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$. In a similar fashion one shows that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$. Hence $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}`$ forms a column-strict tableau of shape $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}`$. We will now show that (8.2) holds with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}`$ as defined in (8.20). First assume that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. In this case $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ if $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}`$. Hence it needs to be shown that in this case $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. To this end it suffices to show that there exists an index $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$ such that (8.22) $$\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ Assume that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$, so that $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. By (8.6) $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}`$ so that (8.22) holds with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. Next assume that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. Then by (8.19), $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}`$. Furthermore by (8.6), $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ which implies (8.22) with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$. In summary (8.22) holds for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. It remains to show that for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ the bounds (8.9) imply (8.2) with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}`$ as in (8.20). First assume $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}`$. For $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$ such that $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}`$ with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ (8.5) simply reads $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. By construction there are no singular strings of length $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{J}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$. Hence, for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$ we can sharpen the bounds in (8.5) and therefore also those in (8.9) by adding the terms (8.23) $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{min}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}$$ if $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ and (8.24) $$\begin{array}{c}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\hfill \\ \hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{min}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\end{array}$$ if $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$. In terms of the paths, this corresponds to adding a horizontal line segment (which is equivalent to extra minus signs) in the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}`$-th strip in the interval $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ whenever there is a horizontal gap between two neighboring paths. An example is given in Figure 2. It depicts the $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}`$-th strip and the zigzag lines correspond to the added line segments. The sum of extra terms (8.23) or (8.24) and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ does not exceed (8.25) $$\begin{array}{c}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{max}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\hfill \\ \hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}\end{array}$$ To obtain the first line of (8.25) we have used $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}`$ by (8.6) for the term $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$, the definition (8.19) of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ when $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ which follows from (8.6). When $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ the second line follows directly using (8.20). When $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ use that $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ by (8.6) so that the last term vanishes. Similarly for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ the bounds in (8.9) can be sharpened by adding $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ Together with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ this yields by similar reasons as before (8.26) $$\begin{array}{c}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\hfill \\ \hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}\end{array}$$ Note that $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}`$. In addition $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}`$. For $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ this follows from (8.7) and Remark 8.10, and for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ this follows from (8.20) and for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ one exploits the first condition of (8.19) and (8.21). Also $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}`$ thanks to (8.6). This implies for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}`$ (8.27) $$\begin{array}{c}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\hfill \\ \hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}\end{array}$$ From (8.25) (or (8.26) for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$) and (8.27) it is straightforward to see that (8.9) implies (8.2) for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}`$. Now consider $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}`$. Then $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}}`$ as shown above (8.27) and $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ so that $$\begin{array}{cc}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\hfill \\ \hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\hfill \\ \hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}\hfill \end{array}$$ For $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ we have $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$ and $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}`$ so that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. For $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ we have $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$ and $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ so that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. This concludes the proof that (8.9) implies (8.2) for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$. ### Case $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$. In this case $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$, $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$, $`\overline{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ and $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}`$ is a tableau over the alphabet $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}`$. It follows from (8.8) that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. Note that this requires in particular that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\overline{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}`$. Define $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$. Then by the column-strictness of $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}`$ we have $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$. Note in particular that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$ so that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}`$ is a column-strict tableau over the alphabet $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}`$. In addition it follows from (8.6) that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\overline{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}`$ so that (8.22) holds for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. This ensures (8.2) for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. Using the fact that there are no singular strings of length $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\overline{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}`$ in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{J}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ and that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\overline{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ by (8.6) the term $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ in (8.9) can be safely replaced by $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\overline{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Furthermore dropping the term $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ equation (8.9) becomes $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ Using $`\overline{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ and the definition of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}`$ this is exactly (8.2). This concludes the proof of the forward direction of the theorem. ### Proof of the reverse direction Let us now prove the reverse direction. To this end consider a rigged configuration $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ corresponding to a column-strict tableau $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}`$ of shape $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}`$ and content $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}`$ under $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}`$ which satisfies $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$ for all $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ and (8.2). We need to show that then $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}`$ is of level $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$. This is equivalent to showing that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ is of level $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$ and that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$ if $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$, where $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}`$ is the row index of the cell $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}`$, $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{shape}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{shape}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. By induction the statement that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ is of level $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$ is equivalent to the statement that $`\overline{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{J}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ satisfies $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$ for all $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ and (8.28) $$\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}$$ for some column-strict tableau $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}`$ of shape $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}`$. To prove $`\overline{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ it suffices to show that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ cannot occur when $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ ($`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}`$) be the length of the selected singular string in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ under $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}}`$. By definition $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$, and the rigged configuration $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{J}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ is obtained from $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ by shortening the selected strings by one, making them singular again and leaving all other strings unchanged. Since $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$ this immediately implies $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$ for all $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$. The vacancy numbers are related by (8.29) $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ Furthermore $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ and $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}`$ so that (8.2) implies (8.30) $$\begin{array}{c}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\hfill \\ \hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\end{array}$$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. Since $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ is the length of a singular string in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ it must be in one of the intervals in (8.4). Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}`$ be the index such that (8.31) $$\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}$$ where recall that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ for all $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$. By similar arguments as in the derivation of (8.8) one finds (8.32) $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ ### Case $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. In this case $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$, $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$, $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\overline{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}`$. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ ($`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$) be maximal such that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$ for all $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$. It follows from Remark 8.10 and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$. Set $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ and $$\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\begin{array}{cc}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\hfill & \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\text{for }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\text{,}}$}\hfill \\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\hfill & \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\text{for }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\text{,}}$}\hfill \end{array}$$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$. The definition of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ and column-strictness of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}`$ ensure the column-strictness of $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}`$. Note that the terms $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ in the two sums in (8.30) cancel each other. Recall that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. Assume $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. The term $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ in (8.30) can be replaced by $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. If $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ this follows from the fact that by construction there are no singular strings of length $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}`$ in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$. For $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ we have $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Hence using $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\overline{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ $$\begin{array}{cc}\hfill \overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\hfill \\ & \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\hfill \\ \hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\hfill \end{array}$$ which is (8.28) for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. Now assume $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$. Since $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ the terms involving $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ in (8.30) vanish and $$\begin{array}{cc}\hfill \overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\hfill \\ \hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\hfill \end{array}$$ which is (8.28) for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$. This concludes the proof that (8.2) implies (8.28) for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. ### Case $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$. Here $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}`$, $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ and $`\overline{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}`$. Set $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ and (8.33) $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\begin{array}{cc}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\hfill & \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\text{for }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\text{,}}$}\hfill \\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{min}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}\hfill & \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\text{for }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\text{,}}$}\hfill \end{array}$$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}`$ where recall that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$. Note that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ or $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ due to (8.32). Define $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$, (8.34) $$\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\begin{array}{cc}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\hfill & \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\text{for }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\text{ and }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\text{,}}$}\hfill \\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{min}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}\hfill & \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\text{for }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\text{,}}$}\hfill \\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\hfill & \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\text{for }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\text{,}}$}\hfill \end{array}$$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}`$ and $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$. Recall that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$. It needs to be shown that $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}`$ is a tableau over the alphabet $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}`$. Since $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}`$ the only problematic case is the third case in (8.34). Condition 1 of the theorem implies that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}`$ so that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$. By (8.31) the condition $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}`$ can only be violated if $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. Assume that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ for some $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}`$. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ be maximal such that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ for all $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$. Then $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ for all $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ so that the second case in (8.33) applies. If $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ we have $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ by the maximality of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$. If $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. In both cases it follows that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ for all $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$. Hence by (8.34) the case $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ does not occur. This proves that $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}`$. It remains to show that $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}`$ is column-strict. The condition $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}`$ only needs to be considered for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}`$ and for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}`$ by the column-strictness of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}`$. In these cases $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}`$ can be deduced from the following inequalities: $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{min}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{min}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{min}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{min}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\text{for }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{min}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\text{for }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}`$ where (8.31) was employed extensively. The condition $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ needs to be verified for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$, for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ and for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$. In these cases $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ can be deduced from the following inequalities: $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{min}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{min}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}`$ where again (8.31) was employed. In addition for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}`$ we have $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{min}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}`$ if $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}`$. If $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}`$ then $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ is only possible if $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ which implies that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$. However in this case $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$. This proves the column-strictness of $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}`$. By definition $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}`$. Hence we need to check that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. It suffices to show that there exists an index $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$ such that (8.35) $$\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ Assume that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. Then $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ and by (8.31) $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ so that (8.35) holds for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. Now assume $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$. Then by (8.33), $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$. Furthermore by (8.31) $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}`$ so that (8.35) holds for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$. For $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ we need to show that (8.30) implies (8.28). Note that for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}`$ we have $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ since $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{min}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ by (8.31). In addition $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$. For $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ this follows directly from (8.34), and for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ we have $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ by (8.31). Since furthermore $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ by (8.31) we have for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}`$ (8.36) $$\begin{array}{c}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\hfill \\ \hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\end{array}$$ where the last term occurs since $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$. Observe that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}`$ since $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{min}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ by (8.31). Hence using (8.36) and (8.34) we have for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}`$ (8.37) $$\begin{array}{cc}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\hfill \\ \hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}& \underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}\hfill \end{array}$$ Since $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$ by Remark 8.10 and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}`$ equation (8.31) implies that either $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ or $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. Note that in both cases (8.2) reads $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$. Since by construction there are no singular strings of length $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ we can add the term $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ to (8.2) for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. This has the effect that the term $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ in (8.30) for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ can be dropped. Note that for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ we have $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ so that the term $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ in (8.37) is zero. For $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ this term is also zero since $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. Since in addition $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$, this proves that (8.30) implies (8.28) for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. Now assume that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}`$. By construction there are no singular strings of length $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$. Therefore the bounds (8.2) and hence also the bounds in (8.30) can be sharpened by adding $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{max}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}$$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ and $$\begin{array}{c}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{max}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\hfill \\ \hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\end{array}$$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$. Adding these to $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ does not exceed $$\begin{array}{c}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{min}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\hfill \\ \hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}\end{array}$$ Using again that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ this can be combined with the term $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ of (8.30) to yield $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Together with (8.37) this proves that (8.30) implies (8.28) for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}`$. Consider $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}`$. Recall that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}`$ so that (8.2) implies $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}}`$. By construction there are no singular strings of length $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$. Hence for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ the bounds in (8.2) and (8.30) can be sharpened by adding $$\begin{array}{c}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{max}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\hfill \\ \hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\end{array}$$ which added to $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ does not exceed (8.38) $$\begin{array}{cc}& \underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{min}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\hfill \\ \hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}& \underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{min}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\hfill \end{array}$$ where in the last line we used that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{min}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ since by definition $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ in this case. The last line of (8.38) also makes sense for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ since then $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ and $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ by (8.31). The last line of (8.38) combined with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ yields $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ using (8.34). For $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ the term $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ vanishes and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Similarly $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ is zero for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Together these results prove (8.28) for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$. This concludes the proof of the reverse direction of the theorem for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$. ### Case $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$. In this case $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$, $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ and $`\overline{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. Then by (8.32) it follows that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. In particular from (8.31), $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ which yields a contradiction when $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ since by assumption $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. Hence the case $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ cannot occur when $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$. Define $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ and $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ for all $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$. Since $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}`$ it follows that $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}`$. The column-strictness of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}`$ immediately implies the column-strictness of $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}`$. Since $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}`$ and there are no singular strings of length $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ we may drop the term $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ in (8.30). In addition dropping the term $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ (8.30) implies for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ $$\begin{array}{cc}\hfill \overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\hfill \\ \hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}{\overset{\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}{\overset{\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}\hfill \end{array}$$ The terms $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ can be added to both sums since they just cancel each other so that we have (8.28) for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. Finally consider the case $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. We have $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ so that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\overline{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}`$. Since the terms $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ in the two sums cancel, (8.28) for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ reduces to $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$, or equivalently $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ as desired. This concludes the proof that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ is of level $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$. ### Zu guter Letzt It remains to show that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$. Define $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}`$ with corresponding rigged configurations $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$, and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{shape}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ be the largest rigging occurring for the strings of length $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$ in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ and let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$ be the row index of the cell $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}`$. Then $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. Denote the length of the selected string in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ under $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}}`$ by $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$. Let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}`$ be maximal such that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$. If no such $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$ exists set $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$. Then $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$. Hence proving $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$ is equivalent to showing that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}`$. If $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}`$ then also $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}`$. Hence assume that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}`$ with $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. For $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}`$ set $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}`$. We will show by descending induction on $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}`$ that (8.39) $$\begin{array}{c}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\hfill \\ \hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{min}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{min}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\end{array}$$ for all $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ where recall that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ if $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$. Let us first show that (8.39) holds for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}`$. This follows directly from (8.2) using that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{min}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{min}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ thanks to the fact that by assumption $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ and $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$. Now assume (8.39) to be true for some $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}`$. We will prove that then $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ and that (8.39) holds for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ if $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. We claim that (8.40) $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}$$ for all $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ where by definition $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. Assume the opposite, namely let $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\kappa }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ be the smallest index such that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\kappa }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\kappa }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$. By (8.39) there are no singular strings of lengths $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\kappa }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\kappa }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\kappa }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ so that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\kappa }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\kappa }$}}`$. By the minimality of $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\kappa }$}`$ and the fact that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ this implies that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\kappa }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\kappa }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ which is a contradiction. This proves (8.40). Note that similar to Remark 8.11 equation (8.39) can be interpreted in terms of $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}`$ non-intersecting paths which all end at position $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$. In this language the condition (8.40) states that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ is to the right of the $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}`$-th path. Since all paths end at $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$ and there are no parts of length greater than $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$ in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ this implies that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$. More precisely, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}`$ if $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ for all $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$. Let us now prove (8.39) at $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. It follows from (8.29) that $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ By (8.40), $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}`$. Hence $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ so that for $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ $$\begin{array}{c}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{min}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\hfill \\ \hfill \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{min}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\end{array}$$ as desired. When $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ then $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$ so that the above inequality still holds. Since we only consider $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ the term $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ does not contribute by (8.40) and in addition $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}`$ can be replaced by $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ in $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{min}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. This proves (8.39) for $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. Since we have shown that (8.39) implies that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ for $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}`$ it follows that $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}}`$. This concludes the proof of Theorem 8.9.
warning/0001/quant-ph0001062.html
ar5iv
text
# Canonical Pairs, Spatially Confined Motion and the Quantum Time of Arrival Problem ## References
warning/0001/hep-ph0001268.html
ar5iv
text
# 1 Introduction ## 1 Introduction In the framework of the colour dipole model of A.H.Mueller it follows that in the high-colour limit $`N_c\mathrm{}`$ the scattering on a heavy nucleus is exactly described by the sum of fan diagrams constructed of BFKL pomerons, each of them splitting into two . This sum seems to be unitary by itself. It is important that no splitting into three or more pomerons, as introduced in , occurs, although formally they give contributions of the same order. Because of this fact, once the splitting vertex is known, construction of the amplitude for the interaction with the nucleus becomes straightforward, reducing to summing BFKL pomeron fan diagrams. This procedure has been well-known since the times of the old Regge-Gribov theory . In the perurbative QCD a pioneering step was taken also many years ago by L.Gribov, E.Levin and M.Ryskin, who summed fan diagrams in the double log approximation and wrote their well-known non-linear GLR equation . In the framework of the BFKL dynamics the necessary tool for constructing fan diagrams is the corresponding triple pomeron vertex, which was found in the colour diplole approach for $`N_c\mathrm{}`$ by A.H.Mueller and B.Patel and in the $`s`$-channel unitarity approach for any number of colours by J.Bartels and M.Wuesthoff . The equivalence of both results was shown in . The equation for the sum of BFKL fan diagrams with this splitting vertex was written by I.Balitsky in his original operator expansion formalism and by Yu.Kovchegov in the colour dipole framework (with a somewhat unconventional form of the coupling to the target) . Its perturbative solution in the region of small non-linearity (outside the saturation region) was studied in . Asymptotic estimates of the solution were presented in . In the present paper we first rederive the BFKL fan diagram equation by direct summation using the standard form of the pomeron-target coupling. It has a form of a simple (and elegant) evolution equation in rapidity $`y`$ for a wave function $`\varphi (y,q)`$ in the momentum space: $$\frac{\varphi (y,q)}{y}=H\varphi (y,q)\varphi ^2(y,q),$$ (1) where $`H`$ is the BFKL Hamiltonian for the so-called semi-amputated function (a similar form was also obtained in ). In spite of its tantalizing simplicity, Eq. (1) does not seem to allow for an analytical treatment except by perturbative methods, not valid in the most interesting region of strong non-linear effects, or by qualitative asymptotic estimates. The bulk of this paper is correspondingly devoted to its numerical analysis. We numerically study evolution of the wave functuion in rapidity, starting from an appropriately chosen initial function. The results are then used to find the structure function of the nucleus at small $`x=e^y`$ and various virtualities $`Q^2`$. Our results show that for a heavy nucleus the longitudinal part of structure function saturates at $`x0`$ to a universal function $`F_{2L}^{(as)}(Q^2)`$, independent of the nucleus atomic number but strongly dependent on $`Q^2`$ in the whole range of $`Q^210^5`$ (GeV/c)<sup>2</sup> explored. In fact $`F_{2L}^{(as)}(Q^2)`$ is just proportional to $`Q^2`$. This reflects the behaviour of the scattering cross -section of a longitudinally polarized $`q\overline{q}`$ pair on the nucleus: at very small $`x`$ it becomes independent of both the pair size and $`x`$. The latter property indicates that the unitarity is restored and the leading singularity in the complex momentum $`j`$ is reduced to a simple pole at $`j=1`$. Note that the limiting cross-section is in accordance with a picture in which the longitudinally polarized photon with a certain probability splits into colour dipoles, which then scatter on the nucleus as on a black disk. For the leading transverse part of the structure function this probability results infinite. Due to this fact both the cross-section and the structure function continue to grow at $`x0`$ approximately as $`\mathrm{ln}(1/x)`$. These results are in agreement with predictions made in . As to the gluon density, we have found it to have a Gaussian shape as a function of $`\xi =\mathrm{ln}k`$ with a center at $`\xi _0y`$. So with the growth of rapidity it propagates towards higher momenta, practically preserving its form, very much like a soliton wave. Obvious limitations on computing time and memory allow to follow its movement up to momenta not higher than of the order $`10^{10}`$ GeV/c. However we expect that this behaviour persists in the model until arbitrary high values of momenta. As a result, at any fixed value of $`k`$ the density eventually goes to zero as $`y\mathrm{}`$. In this sense we have the maximal saturation of the density possible. The paper is organized as follows. In section 2 we present derivation of the BFKL fan diagram equation by direct summation. Section 3 is devoted to the numerical solution of Eq. (1). In Section 4 the nuclear structure function and gluon density are calculated. Section 5 contains our conclusions. In the Appendix we compare our BFKL fan diagram equation with the Kovchegov’s one . ## 2 Fan diagram equation for the BFKL pomerons We start with a single scattering contribution to the forward ampitude for the interaction of the projectile particle with a nucleus. In the BFKL framework, at fixed impact parameter $`b`$, it has a well-known form $$𝒜_1(y,b)=isg^4AT(b)d^2rd^2r^{}\rho (r)G(y,r,r^{})\rho _N(r^{}).$$ (2) Here $`\rho `$ and $`\rho _N`$ are the colour densities of the projectile and the target nucleon respectively. Function $`G`$ is the forwrad BFKL Green function : $$G(y,r,r^{})=\frac{rr^{}}{32\pi ^2}\underset{n=\mathrm{}}{\overset{+\mathrm{}}{}}e^{in(\varphi \varphi ^{})}_{\mathrm{}}^{\mathrm{}}\frac{d\nu e^{y\omega (\nu )}}{[\nu ^2+(n1)^2/4][\nu ^2+(n+1)^2/4]}(r/r^{})^{2i\nu },$$ (3) where $`\varphi `$ and $`\varphi ^{}`$ are the azimuthal angles and $$\omega (\nu )=2(\alpha _sN/2\pi )(\psi (1)\mathrm{Re}\psi (1/2+i\nu ))$$ (4) are the BFKL levels. Due to the azimuthal symmetry of the projectile colour density one may retain only the term with zero orbital momenta $`n=0`$ in (3). Separating the projectile part, the single scattering term may be written in the form $$𝒜_1(y,b)=2isd^2r\rho (r)\mathrm{\Phi }_1(y,b,r),$$ (5) where $$\mathrm{\Phi }_1(y,b,r)=\frac{1}{2}g^4AT(b)d^2r^{}G(y,r,r^{})\rho _N(r^{}).$$ (6) The double scattering contribution has been calculated in . In the limit $`N_c>>1`$ one finds $$𝒜_2(y,b)=is\frac{g^2N_c}{4\pi ^3}d^2r\rho (r)_0^y𝑑y_2\underset{i=1}{\overset{3}{}}d^2r_i\delta ^2(r_1+r_2+r_3)$$ $$\frac{r_1^2_1^4}{r_2^2r_3^2}G(yy_2,r,r_1)\mathrm{\Phi }_1(y_2,b,r_2)\mathrm{\Phi }_1(y_2,b,r_3).$$ (7) Presenting $`𝒜_2`$ in terms of $`\mathrm{\Phi }_2`$ similarly to (5) we find $$\mathrm{\Phi }_2(y,b,r)=\frac{g^2N_c}{8\pi ^3}_0^y𝑑y_2\underset{i=1}{\overset{3}{}}d^2r_i\delta ^2(r_1+r_2+r_3)$$ $$\frac{r_1^2_1^4}{r_2^2r_3^2}G(yy_2,r,r_1)\mathrm{\Phi }_1(y_2,b,r_2)\mathrm{\Phi }_1(y_2,b,r_3).$$ (8) The whole set of fan diagrams will be evidently summed by the equation which is graphically illustrated in Fig. 1: $$\mathrm{\Phi }(y,b,r)=\mathrm{\Phi }_1(y,b,r)\frac{g^2N_c}{8\pi ^3}_0^y𝑑y_2\underset{i=1}{\overset{3}{}}d^2r_i\delta ^2(r_1+r_2+r_3)$$ $$\frac{r_1^2_1^4}{r_2^2r_3^2}G(yy_2,r,r_1)\mathrm{\Phi }(y_2,b,r_2)\mathrm{\Phi }(y_2,b,r_3).$$ (9) The impact parameter $`b`$ appears here only as a parameter and the dependence on it will be implicit in the following. In terms of $`\mathrm{\Phi }`$ the total forward scattering amplitude on the nucleus will be given at fixed $`b`$ by $$𝒜(y,b)=2isd^2r\rho (r)\mathrm{\Phi }(y,b,r).$$ (10) One can rewrite Eq. (9) as an evolution equation in $`y`$. To this end we present $`\mathrm{\Phi }`$ as an integral over $`\nu `$’s similar to (3): $$\mathrm{\Phi }(y,r)=𝑑\nu r^{12i\nu }\mathrm{\Phi }(y,\nu ).$$ Then from (9) we find $$\mathrm{\Phi }(y,\nu )=\frac{g^4}{64\pi ^2(\nu ^2+1/4)^2}e^{\omega (\nu )y}d^2r^{}r^{1+2i\nu }\rho (r^{})AT(b)$$ $$\frac{g^2N_c}{32\pi ^5(\nu ^2+1/4)^2}e^{\omega (\nu )y}_0^y𝑑y_2e^{\omega (\nu )y_2}\underset{i=1}{\overset{3}{}}d^2r_i\delta ^2(r_1+r_2+r_3)$$ $$\frac{r_1^2_1^4}{r_2^2r_3^2}r_1^{12i\nu }\mathrm{\Phi }(y_2,r_2)\mathrm{\Phi }(y_2,r_3).$$ (11) Multiplying both parts by $`e^{\omega (\nu )y}`$ and taking a derivative in $`y`$ we obtain $$\left(\frac{}{y}\omega (\nu )\right)\mathrm{\Phi }(y,\nu )=\frac{g^2N_c}{32\pi ^5(\nu ^2+1/4)^2}\underset{i=1}{\overset{3}{}}d^2r_i\delta ^2(r_1+r_2+r_3)$$ $$\frac{r_1^2_1^4}{r_2^2r_3^2}r_1^{12i\nu }\mathrm{\Phi }(y,r_2)\mathrm{\Phi }(y,r_3).$$ (12) Returning to the $`r`$ space we take into account that $$\omega (\nu )r^{12i\nu }=\widehat{H}r^{12i\nu },$$ (13) where $`\widehat{H}`$ is the BFKL Hamiltonian . Then (12) transforms into $$\left(\frac{}{y}+\widehat{H}\right)\mathrm{\Phi }(y,r)=\frac{g^2N_c}{8\pi ^3}\underset{i=1}{\overset{3}{}}d^2r_i\delta ^2(r_1+r_2+r_3)\frac{r_1^2_1^4}{r_2^2r_3^2}G(0,r,r_1)\mathrm{\Phi }(y,r_2)\mathrm{\Phi }(y,r_3).$$ (14) Using $$\frac{r_1^2_1^4}{r^2}G(0,r,r_1)=\frac{1}{2\pi ^2rr_1}\underset{n=\mathrm{}}{\overset{+\mathrm{}}{}}e^{in(\varphi \varphi _1)}_{\mathrm{}}^{\mathrm{}}𝑑\nu (r/r_1)^{2i\nu }=\delta ^2(rr_1),$$ (15) we simplify (14) to a form $$\left(\frac{}{y}+\widehat{H}\right)\mathrm{\Phi }(y,r)=\frac{g^2N_c}{8\pi ^3}\underset{i=2}{\overset{3}{}}d^2r_i\delta ^2(r+r_2+r_3)\frac{r^2}{r_2^2r_3^2}\mathrm{\Phi }(y,r_2)\mathrm{\Phi }(y,r_3).$$ (16) The initial condition is determined from (9) to be $$\mathrm{\Phi }(y,r)_{y=0}=\mathrm{\Phi }_0(r)$$ (17) with $$\mathrm{\Phi }_0(r)=\frac{1}{2}g^4AT(b)d^2r^{}G(0,r,r^{})\rho (r^{}).$$ (18) Except for the initial condition and different variables, Eq. (16) coincides with the one constructed by Kovchegov in the colour dipole approach in (see Appendix). Now we go to function $$\varphi (y,r)=\frac{1}{r^2}\mathrm{\Phi }(y,r)$$ (19) and pass to the momentum space. For $`\varphi (y,q)`$ Eq. (16) reads $$\left(\frac{}{y}+\frac{1}{r^2}\widehat{H}r^2\right)\varphi (y,q)=\frac{g^2N_c}{8\pi ^3}\varphi (y,q)\varphi (y,q),$$ (20) where we have used that $`\varphi (y,q)`$ in fact depends only on $`|q|`$, which follows from the initial conditon $$\varphi _0(y,q)_{y=0}=\frac{d^2r}{r^2}e^{iqr}\mathrm{\Phi }_0(r)$$ (21) with $`\mathrm{\Phi }_0`$ given by (18) and depending only on $`|r|`$ due to the azimuthal summetry. The Hamiltonial $`r^2\widehat{H}r^2`$ which appears on the left-hand side is the standard forward Hamiltonian for the semi-amputated functions $$\frac{1}{r^2}\widehat{H}r^2=H=\frac{g^2N_c}{4\pi ^2}[\mathrm{ln}q^2+\mathrm{ln}r^22(\mathrm{ln}2+\psi (1))].$$ (22) Indeed the eigenfunctions of $`\widehat{H}`$ $`\mathrm{\Phi }_{n,\nu }(r)=r^{12i\nu }e^{in\varphi }`$ go over into the eigenfunctions of $`H`$ after division by $`r^2`$. This brings us to nearly the final form of our equation $$\left(\frac{}{y}+H\right)\varphi (y,q)=\frac{g^2N_c}{8\pi ^3}\varphi ^2(y,q).$$ (23) It remains only to appropriately rescale $`H`$, $`\varphi `$ and $`y`$ to obtain the BFKL fan diagram equation in the final form (1). We introduce $$H=\frac{g^2N_c}{4\pi ^2}\stackrel{~}{H},\stackrel{~}{y}=\frac{g^2N_c}{4\pi ^2}y,\mathrm{and}\varphi (\stackrel{~}{y},q)=2\pi \stackrel{~}{\varphi }(\stackrel{~}{y},q),$$ (24) where $$\stackrel{~}{H}=\mathrm{ln}r^2+\mathrm{ln}q^22(\mathrm{ln}2+\psi (1)).$$ (25) The equation for $`\stackrel{~}{\varphi }`$ takes the form (1) $$\left(\frac{}{d\stackrel{~}{y}}+\stackrel{~}{H}_0\right)\stackrel{~}{\varphi }(\stackrel{~}{y},q)=\stackrel{~}{\varphi }^2(\stackrel{~}{y},q)$$ (26) with the initial condition $$\stackrel{~}{\varphi }(\stackrel{~}{y},q)_{\stackrel{~}{y}=0}=\frac{g^4AT(b)}{4\pi }d^2rd^2r_1\frac{1}{r^2}e^{iqr}G(0,r,r_1)\rho _N(r_1).$$ (27) ## 3 Numerical solution of the BFKL fan diagram equation To solve numerically Eq. (26) we first transform the BFKL Hamiltonian $`H`$ to more convenient variables. In the momentum space one can write action of $`H`$ on a function $`\varphi (q)`$ as $$H\varphi (q)=2_0^{\mathrm{}}k𝑑k\left[\frac{\varphi (k)(q^2/k^2)\varphi (q)}{|q^2k^2|}+\frac{q^2}{k^2}\frac{\varphi (q)}{\sqrt{4k^4+q^4}}\right].$$ (28) We transform it to variables $`u,v`$ which take values in : $$q=\mathrm{exp}[(M_1+M_2)uM_1],k=\mathrm{exp}[(M_1+M_2)vM_1],$$ (29) where $`q,k`$ are in GeV/c and $`M_{1(2)}`$ is a lower(upper) integration limit in $`\mathrm{ln}k`$. Then (28) goes into $$H\varphi (u)=2(M_1+M_2)_0^1𝑑v\left[\frac{\varphi (v)f(uv)\varphi (u)}{|f(uv)1|}+\frac{\varphi (u)}{\sqrt{1+4f^2(uv)}}\right],$$ (30) where $$f(u)=\mathrm{exp}[(M_1+M_2)u].$$ (31) Now we discretize the interval in $`u`$ and $`v`$ into $`n`$ equidistant points $`u_0,u_1,\mathrm{}u_n`$ and $`v_0,v_1,\mathrm{}v_n`$ to convert the evolution equation (26) into a set of $`n`$ ordinary 1st order non-linear differential equations in $`y`$. This set of equations can be solved by standard methods. We used the simplest 2nd order Runge-Kutta algorithm. The standard values of $`n`$ and of the number of iterations to evolve in two units of $`\stackrel{~}{y}`$ were 800. We have studied evolution of $`\varphi (y,q)`$ from the initial value at $`y=0`$ up to $`\stackrel{~}{y}=10`$. The (fixed) value of the strong coupling $`\alpha _s=g^2/4\pi `$ has been chosen to be 0.2. With this choice our maximal rapidity is around 50. An evident difficulty which one meets are the values of $`H\varphi `$ at endpoints $`u_0`$ and $`u_n`$, at which the introduced cutoffs make the results not reliable. To overcome this difficulty we calculated $`H\varphi `$ at these points by extrapolation from the neighbouring points. The stability of this procedure was checked by comparing the results for double and quadruple values of $`n`$. The initial function is determined by the colour density of the nucleon $`\rho _N(r)`$ according to (27). To simplify our calculations we have taken the Yukawa form for $`\rho _N(r)`$: $$\rho _N(r)=\frac{\mu }{2\pi }\frac{e^{\mu r}}{r},$$ (32) where $`\mu `$=1/0.7 fm has a meaning of the inverse nucleon radius. This choice may look a bit arbitrary, but our results show that with the growing rapidity the system quickly forgets not only the form but even the absolute magnitude of $`\rho _N`$, so that the solution becomes independent of the initial function and governed only by the internal dynamics of Eq. (26) itself. Doing the integrations over $`r`$ and $`r_1`$ in (27) and over $`\nu `$ inside the BFKL Green function (3) we find for $`q>\mu `$ $$\stackrel{~}{\varphi }_0(q)_{q>\mu }=B\frac{\mu ^2}{q^2}\left\{1\frac{\mu }{q\sqrt{\pi }}\underset{n=0}{}(1)^n\left(\frac{\mu ^2}{q^2}\right)^n\frac{\mathrm{\Gamma }(1/2+n)}{(n+3/2)^2}\right\}$$ (33) and for $`q<\mu `$ $$\stackrel{~}{\varphi }_0(q)_{q<\mu }=B\{\frac{1}{4}[(2\mathrm{ln}\frac{q}{\mu }+\psi (3/2)\psi (1)1)^2+\psi ^{}(3/2)+\psi ^{}(1)+1]$$ $$\frac{q^2}{\sqrt{\pi }\mu ^2}\underset{n=0}{}(1)^n\left(\frac{q^2}{\mu ^2}\right)^n\frac{\mathrm{\Gamma }(5/2+n)}{(n+2)(n+1)^3}\},$$ (34) where the dimensionless coefficient $$B=\frac{g^4}{16\pi }\frac{AT(b)}{\mu ^2}$$ (35) carries all the information about the nucleus. Its maximal value with $`\alpha _s=0.2`$ is about 0.12 for the central scattering on lead ($`b=0`$). The asymptotic behaviour of $`\varphi _0(q)`$ at $`q\mathrm{}`$ and $`q0`$ is governed by the first terms in (33) and (34), which come from the poles of the BFKL Green functions at $`\nu =\pm i/2`$: $$\varphi _0(q)1/q^2,q\mathrm{},$$ $$\varphi _0(q)\mathrm{log}^2q,q0.$$ (36) We have checked that our results parctically do not change if the complicated function which multiplies $`B`$ in (33) and (38) is substituted just by $`\mathrm{ln}^2(q/(\mu +q)`$ with the same asymptotic behaviour (36). The initial function depends on the form of the nucleus profile function $`T(b)`$. For our numerical calculation we have chosen $$T(b)=2\sqrt{R_A^2b^2}/V_A$$ (37) corresponding to a finite nucleus of radius $`R_A`$ and volume $`V_A`$ with a constant density. We have taken $`R_A=A^{1/3}R_0`$ with $`R_0=1.2`$ fm. Our results for $`\varphi (q)`$ for two values of $`B=0.12`$ and 0.02, corresponding to a central and very peripheral collisions off lead, respectively, are presented in Figs 2,3. One observes that the solution $`\varphi `$ at rapidities $`\stackrel{~}{y}>2`$ evolves to a very simple and universal form, practically independent of the initial function. Crudely speaking it linearly falls with $`\xi =\mathrm{ln}k`$ until it meets the $`x`$-axis wherefrom it stays equal to zero. The slope of the falling part is exactly equal to 1, so that very crudely $$\varphi (\xi )=\xi _1(y,b)\xi ,\mathrm{for}\xi <\xi _1(y,b),\varphi (\xi )=0,\mathrm{for}\xi >\xi _1(y,b).$$ (38) The value of $`\xi _1(y,b)`$ and hence the interception point with the $`x`$ axis grow linearly with $`y`$, so that with the growth of $`y`$ the picture simply shifts to the right. In reality the curve for $`\varphi `$ of course has no break: the two straight lines of which it is formed join smoothly in the vicinity of $`\xi _1`$. As will be clear later, the physically important region is precisely this vicinity, where $`\varphi `$ is not trivial. The results for $`\varphi (k)`$ do not depend on the chosen cutoffs, provided they are taken to cover the region around the interception point $`\xi _1`$. Otherwise the evolution stops as soon as $`\xi _1`$ touches the upper cutoff $`M_2`$. So if one wants to study evolution up to high values of $`y`$ the upper cutoff should be taken correspondingly high. In our calculations we chose $`M_1=10`$ and $`M_2=20`$, having verified that further raising of either $`M_1`$ or $`M_2`$ does not change the results. In conclusion we find that $`\varphi `$ does not possess any finite limit as $`y\mathrm{}`$ so that Eq. (26) does not lead to any saturation of the wave function $`\varphi `$ at high rapidities, contrary to naive expectations. However in the next chapter we shall see, that such saturation indeed occurs for physical quantities. The point is that function $`\varphi `$ by itself has no physical meaning. It is its derivatives which matter. ## 4 Nuclear structure function and the gluon density The nuclear structure function is obtained in the standard manner as $$F_2(x,Q^2)=\frac{Q^2}{\pi e^2}(\sigma _T+\sigma _L),$$ (39) where $`\sigma _{T,L}`$ are the total cross-section for the scattering on the nucleus of a virtual photon with transversal (T) or longidunal (L) polarization. Both cross-sections can be found from the imaginary part of the forward scattering amplitude (10). In terms of $`\stackrel{~}{\varphi }`$ we find $$\sigma _{T,L}=4\pi d^2bd^2r\rho _{T,L}(r)r^2\stackrel{~}{\varphi }(y,r,b).$$ (40) Here we explicitly indicated the dependence of $`\varphi `$ on the impact parameter; $`\rho _{T,L}(r)`$ are the well-known colour densities of the virtual photon split into a $`q\overline{q}`$ pair (see e.g. ). With massless quarks $$\rho _T(r)=\frac{e^2N_cZ^2}{8\pi ^3}_0^1d\alpha (\alpha ^2+(1\alpha )^2)ϵ^2\mathrm{K}_1^2(ϵr))$$ (41) and $$\rho _L(r)=\frac{e^2N_cZ^2}{2\pi ^3}Q^2_0^1𝑑\alpha \alpha ^2(1\alpha )^2\mathrm{K}_0^2(ϵr),$$ (42) where $`ϵ^2=Q^2\alpha (1\alpha )`$ and $`Z^2`$ is a sum of squares of quark electric charges in units $`e`$. Passing to momentum space we find $$\sigma _{T,L}=4\pi d^2b\frac{d^2q}{(2\pi )^2}\stackrel{~}{\varphi }(q,y,b)w_{T,L}(q),$$ (43) where $$w_{T,L}(q)=d^2rr^2\rho _{T,L}(r)e^{iqr}.$$ (44) Straightforward calculation leads to the following expressions for $`w_{T,L}(q)`$. For the transverse density one finds<sup>1</sup><sup>1</sup>1An error in this formula lead to some erroneous conclusuions about the $`y`$-behaviour of the structure function in the original version $$w_T(q)=\frac{e^2N_cZ^2}{8\pi ^2}_0^1𝑑\alpha (\alpha ^2+(1\alpha )^2)_q^2[(q^2/2+ϵ^2)J(q,ϵ)],$$ (45) where $$J(q,ϵ)=\frac{2}{q\sqrt{q^2+4ϵ^2}}\mathrm{ln}\frac{\sqrt{q^2+4ϵ^2}+q}{\sqrt{q^2+4ϵ^2}q}.$$ (46) The longitudinal density $`w_L`$ is given by the same expression with substitutions $$\alpha ^2+(1\alpha )^2\alpha (1\alpha ),q^2/2+ϵ^24ϵ^2$$ With the found numerical values for the function $`\varphi `$ in the range $`0\stackrel{~}{y}10`$ we evaluated the nuclear structure function of lead ($`A=207`$) for various $`Q^2`$ between 3 and 10<sup>5</sup> (GeV/c)<sup>2</sup>. The results are most instructive for the the cross-section $`\sigma _L`$ for the scattering of a longitudinally polarized virtual photon on the nucleus. In Figs. 4,5 we present it (with $`e^21`$) as a function of $`y`$ at fixed $`Q^2`$ and as a function of $`Q^2`$ at fixed $`y`$ respectively (in GeV<sup>-2</sup>). From Fig. 4 one clearly sees saturation in rapidity: for any value of $`Q^2`$ the cross-section tends to the same limit of $`228.8`$ (GeV/c)$`{}_{}{}^{2}=0.1768R_A^2`$ as $`\stackrel{~}{y}`$ goes beyond 5. The resulting constant cross-section is evidently consistent with the unitarity restrictions. In terms of the complex angular momentum $`j`$ our results indicate that the original cut at $`j>1`$ is reduced to a simple pole at $`j=1`$. Fig. 5 illustrates an unusual behaviour in $`Q^2`$ which sets in at high $`y`$: instead of going down as $`1/Q`$ in the standard BFKL approach, the cross-section becomes independent of $`Q^2`$ to a very high precision. These results for the longitudinal cross-section can be conveniently interpreted in terms of scattering of colour dipoles off the nucleus. The density $`\rho _L(r)`$, appropriately normalized, can be interpreted as a probability distribution for the longitudinal photon to split into colour dipoles of transverse dimension $`r`$. The normalization factor $$D=d^2r\rho _L(r)=\frac{e^2N_cZ^2}{12\pi ^2}=0.028145e^2$$ can be considered as a total probability for the photon to split into dipoles. Then the dipole-nucleus cross-section is found by dividing $`\sigma _L`$ by factor $`D`$, which gives 8131.1 (GeV/c)<sup>-2</sup> independent of $`Q^2`$, that is, of the dipole dimension. This value exactly equals $`2\pi R_A^2`$, corresponding to scattering off a black disk. The transverse part of the structure function does not admit this interpretation, due to the fact that $`\rho _T(r)`$ is not normalizable. As a result both the cross-section and the structure function do not saturate at large $`y`$ but continue to grow nearly linearly in $`y`$. This is illustrated in Figs. 6,7 where the total structure function (including the much smaller longitudinal part) is shown as a function of $`x=e^y`$ and $`Q^2`$ respectively. We finally come to the gluon density. Although, strictly speaking it is not a physical quantity, its properies have been much discussed recently in connection with its saturation for the large nucleus . It can be related to our function $`\varphi `$ via the standard expression for the structure function in its terms (see e.g ) $$F_2(x,Q^2)=\frac{g^2Q^2}{\pi ^3Ne^2}d^2bd^2r[\rho _T(r)+\rho _L(r)]F(x,r,b),$$ (47) where $$F(x,r,b)=\frac{d^2k}{(2\pi )^2k^4}k^2\frac{xG(x,k^2,b)}{k^2}\left(1e^{ikr}\right)\left(1e^{ikr}\right)$$ (48) and $`d^2b(xG(x,k^2,b)/k^2)`$ is up to a factor the gluon density in the momentum space: $$\frac{N(l)}{^2l}=\frac{1}{\pi }\frac{N(l)}{l^2}=\frac{1}{\pi }d^2b\frac{xG(x,l^2,b)}{l^2}.$$ (49) In the following we shall study the double density in momentum and impact parameter, which is just $`xG(x,k^2,b)/k^2`$. Comparing (47), (48) with the corresponding expression in terms of $`\varphi `$, which follows from (39) and (40) we find a relation $$F(x,r,b)=\frac{4N}{\pi g^2}r^2\stackrel{~}{\varphi }(\mathrm{ln}\frac{1}{x},r,b).$$ (50) Taking a Fourier transform of (48) and neglecting the term proportional to $`\delta ^2(k)`$ we obtain $$\frac{xG(x,k^2,b)}{k^2}=\frac{2N}{\pi g^2}k^2_k^2\stackrel{~}{\varphi }(\mathrm{ln}\frac{1}{x},k,b).$$ (51) This is the desired relation between our function $`\varphi `$ and the gluon density in the combined momentum and impact parameter space. Applying $`k^2_k^2`$ to the found function $`\varphi `$ we thus find the gluon density up to a trivial numerical factor evident from (51)( $`0.76`$ with $`\alpha _s=0.2`$). Function $`h(k)=k^2_k^2\stackrel{~}{\varphi }(y,k,b)`$ for different values of $`y`$ and $`B=0.12`$ and 0.02 is shown in Figs. 8,9. Its form at different $`y`$ results quite remarkable. As one observes, at any given rapidity the found density has the same, roughly Gaussian shape in variable $`\xi =\mathrm{ln}k`$, centered at the point $`\xi =\xi _0(y)`$ very near to $`\xi _1`$ at which the straight line (38) crosses the $`x`$-axis (see Section 2). With the growth of $`y`$ the distribution moves to the right with a nearly constant velocity practically preserving its form. Approximately the distribution can be described by $$h(k)=h_0e^{a(\xi \xi _0(y))^2},$$ (52) where $`h_0`$ and $`a`$ are practically independent of $`y`$ and $`\xi _0(y)`$ linearly grows with it: $$h_00.3,a0.3\xi _0(y)=\xi _{00}+2.23\stackrel{~}{y}.$$ (53) The only quantity which clearly depends on the initial distribution is the starting position $`\xi _{00}`$, so that for different initial functions the picture in Figs. 8 and 9 shifts along the $`\xi `$ axis as a whole. Evidently at a given value of $`k`$ the density stays always limited, irrespective of the form of the initial distribution (and on the atomic number $`A`$, in particular). In this sense we have saturation as discussed in . However with the growth of $`y`$ the strongly peaked density moves away toward higher values of $`k`$ so that the density at a fixed point tends to zero at high values of rapidity. We thus have “supersaturation”: with $`y\mathrm{}`$ the gluon density at an arbitrary finite momentum tends to zero. Comparing Figs. 8 and 9 one can see how the memory about the initial distribution (except for $`\xi _{00}`$) is gradually erased in the course of the evolution in $`y`$. For a very peripheric collison off lead ($`B=0.02`$) at the initial stages of the evolution the density is correspondingly much smaller than for a central collision ($`B=0.12`$). However already at $`\stackrel{~}{y}=2`$ the form of the distribution is practically indistinguishable from the central collision, the only remaining difference being the shift along the $`\xi `$ axis. ## 5 Conclusions The BFKL fan diagram equation has been solved numerically in the large range of rapidities up to $`y=50`$. The main results are the following. The idea that the fan diagrams themselves satisfy the unitarity condition has been supported by the fact that the found cross-sections for the scattering of a $`q\overline{q}`$ pair off the nucleus tend to a constant value at high rapidities. Since the found cross-sections do not rise with energy, the leading $`j`$-plane singularity turns out to be a simple pole at $`j=1`$. The limiting cross-sections prove to be universal: they do not depend on $`Q^2`$, that is, on the transverse dimension of the $`q\overline{q}`$ pair, nor on the initial colour distribution inside the nucleon. Their dependence on the target nucleus thus reduces to a scale factor $`R_A^2`$. Physically they correspond to scattering of a colour dipole off a black disk. The nuclear structure function does not saturate at high rapidities, due to the singularity of the transverse distribution $`\rho _T(r)`$ at $`r=0`$, which makes it non-normalizable. At large $`y`$ it continues to grow nearly linearly in $`y`$. These results fully agree with predictions made in on the basis of the found perturbative solution of the BFKL fan diagram equation and asymptotic estimates made in . A completely novel result concerns the gluon density of the nucleus. At sufficiently high rapidities, greater than 10, the gluon density aquires a form of the soliton wave in $`y\mathrm{ln}k`$ space, which, with the growth of $`y`$, moves along the $`\mathrm{ln}k`$ towards greater $`k`$ preserving its nearly Gaussian shape. Thus at any finite $`k`$ the gluon density eventually goes to zero at high enough $`k`$ Finally we have to stress that all these properties begin to be clearly visible only at very high rapidities and momenta: $`y>10`$ and $`k>100`$ GeV/c with $`\alpha _s=0.2`$. With smaller $`\alpha `$’s these values grow correspondingly. ## 6 Acknowledgements The author is grateful to Dr.G.P.Vacca who draw his attention to a possibility to simplify the BFKL fan diagram equation. He is also grateful to Prof. Bo Andersson for his interest in this work and stimulating discussions. He thanks the Department for Theoretical Physics of the Lund University for hospitality during his stay in Lund, where this work was completed. The author is finally most thankful to Dr. Yu.Kovchegov, fruitful discussions with whom helped to find and correct an error in the expression for $`w_T`$ (Eq. (45)) in the first version of this paper. ## 7 Appendix. Colour dipole approach To compare our BFKL fan diagram equation to that of Kovchegov we present here a short derivation of Eq. (1) from the colour dipole approach, which will make clear the difference between the two equations. In the colour dipole approach the single scattering term (2) is presented in the form $$𝒜(y,b)=2isAT(b)d^2r\rho (r)d^2r_1n_1(r,r_1,y)\tau (r_1),$$ (54) where $$\tau (r_1)=\frac{1}{2}g^4d^2r^{}G(0,r_1,r^{})\rho _N(r^{})$$ (55) and $`n_1(r,r_1,y)`$ is a single dipole density at rapidity $`y`$ introduced by A.H.Mueller; $`r_1`$ and $`r`$ are the dipole lengths at rapidity $`y`$ and at $`y=0`$ respectively, so that $$n_1(r,r_1,y)_{y=0}=\delta ^2(rr_1).$$ (56) (Note that as in our $`n_1`$ is A.Mueller’s one divided by $`2\pi r_1^2`$.) To introduce the multidipole densities A.H.Muller constructed a generating functional $`Z(r_1,r_0,y|u)`$, where $`u=u(\rho _i,\rho _f)`$ is a function of two dipole endpoints in the transverse space (which, for brevity, Iwe denote by a single symbol $`\rho `$ for). The functional $`Z`$ satisfies the following nonlinear equation $$Z(r_1,r_2,y|u)=u(r_1,r_0)e^{2y\omega (r_{10})}+$$ $$\frac{g^2N_c}{8\pi ^3}_0^y𝑑y^{}e^{2\omega (r_{10})(yy^{})}d^2r_2\frac{r_{10}^2}{r_{12}^2r_{20}^2}Z(r_1,r_2,y^{}|u)Z(r_2,r_0,y^{}|u),$$ (57) where $`r_{10}=r_1r_0`$ etc and $`\omega (r)`$ is the gluon trajectory $`\omega (q)`$ in which the momentum $`q`$ is substituted by $`r`$: $$\omega (r)=\frac{g^2N_c}{4\pi ^3}\mathrm{ln}\frac{r}{ϵ}$$ (58) with $`ϵ`$ the cutoff at small $`r`$ (in the ultraviolet). The functional $`Z`$ is normalized according to $$Z(r_1,r_0,y|u)_{u=1}=1.$$ (59) The $`k`$-fold inclusive dipole density is given by a $`k`$-fold derivative of $`D`$ with respect to $`u`$ at $`u=1`$: $$n_k(r_1,r_0,y;\rho _1,\mathrm{}\rho _k)=\frac{1}{k!}\frac{\delta ^kZ}{\delta u(\rho _1)\mathrm{}.\delta u(\rho _k)}_{u=1}.$$ (60) At this point we make our first comment as to the comparison with . Our form of the functional equation for $`Z`$ is the same as in the original pater of A.H.Mueller and in , except for a slightly different choice of dipole coordinates and for the order of arguments in the 2nd $`Z`$ in the non-linear term: their form would correspond to $`Z(r_0,r_2,y^{}|u)`$. Comparing with the BFKL equation for the single dipole density, one can verify that our choice is better. However this point is irrelevant for the following, since in the interaction with the nucleus only even functions of $`r_{10}`$ appear. The dipoles are to interact with the nucleus target with a zero transferred momentum. If the dimension of the dipole is smaller than the internucleon distance in the nucleus then it will interact with a single nucleon as a whole. This picture lies at the basis of the standard fan diagram approach, where each pomeron finally interacts with a single nucleon. We used precisely this picture in our derivation of Eq. (26) in Section 2. In this picture the only trace of the nucleus will be an additional factor $`AT(b)`$ multiplying the interaction with the nucleon $`\tau (r)`$ (Eq. (54)) at a given impact parameter $`b`$. In a different idea is exploited: it is assumed that each of the dipoles can interact with many nucleons. The latter interaction is assumed to have an eikonal form. So the interaction with the nucleus they consider is a two-stage one: first the projectile generates many colour dipoles (BFKL fan diagrams) and then each dipole multiply interacts with the nucleus a-la Glauber. Although technically it is not difficult to take into account the final eikonalization of the interaction, appropriately changing the single scattering term $`\tau (r)`$ in (55) and the following formulas, we do not think it is reasonable. On the one hand, as we shall see, in the interaction with the nucleus the densities are considerably damped at large distances as compared to the usual BFKL behaviour. The confinement should further restrict their spatial dimensions. So for a nucleus with a large internucleon distance it does not seem reasonable to assume simultaneous interaction of a dipole with two or more nucleons. On the other hand, should such interactions be really important, one cannot expect to correctly describe the interaction with the nucleus of a dipole of a given (and fixed) dimension by the Glauber formula. Note that the expression used in for it does not correspond to the BFKL picture for the scattering on a single nucleon. So we take $`\tau `$ as given by (55) and this is the main difference between our derivation and that of . With a chosen $`\tau `$, the interaction with the nucleus will be described by densities $$\nu _k(r_1,r_0,y)=\underset{j=1}{\overset{k}{}}\left(d^4\rho _j\tau (\rho _j)AT(b)\right)n_k(r_1,r_0,y;\rho _1,\mathrm{}\rho _k).$$ (61) Here $`\tau (\rho )=\tau (\rho _f\rho _i)`$ depends only on the dipole length and is given by (55). Differentiating (57) and using (61) one easily obtains $$\nu _1(r_1,r_0,y)=AT(b)\tau (r_{10})e^{2y\omega (r_{10})}+$$ $$\frac{g^2N_c}{8\pi ^3}_0^y𝑑y^{}e^{2\omega (r_{10})(yy^{})}d^2r_2\frac{r_{10}^2}{r_{12}^2r_{20}^2}[\nu _1(r_1,r_2,y^{})+\nu _1(r_2,r_0,y^{})]$$ (62) and for $`k>1`$ $$\nu _k(r_1,r_0,y)=\frac{g^2N_c}{8\pi ^3}_0^ydy^{}e^{2\omega (r_{10})(yy^{})}d^2r_2\frac{r_{10}^2}{r_{12}^2r_{20}^2}[\nu _k(r_1,r_2,y^{})+\nu _k(r_2,r_0,y^{})$$ $$+\underset{j=1}{\overset{k1}{}}\nu _j(r_1,r_2,y^{})\nu _{kj}(r_2,r_0,y^{})].$$ (63) We suppress the evident dependence on the impact parameter $`b`$. From the structure of the equations and the form of the inhomogeneous term it follows that the densities $`\nu (r_1,r_0,y)`$ depend only on the initial dipole length $`r_{10}`$. As a result, the two terms separated from the sum over $`j`$ on the right-hand side give the same contribution. One then finally finds equations $$\nu _1(r_{10},y)=AT(b)\tau (r_{10})e^{2y\omega (r_{10})}+\frac{g^2N_c}{4\pi ^3}_0^y𝑑y^{}e^{2\omega (r_{10})(yy^{})}d^2r_2\frac{r_{10}^2}{r_{12}^2r_{20}^2}\nu _1(r_{20},y^{})$$ (64) and for $`k>1`$ $$\nu _k(r_{10},y)=\frac{g^2N_c}{4\pi ^3}_0^y𝑑y^{}e^{2\omega (r_{10})(yy^{})}d^2r_2\frac{r_{10}^2}{r_{12}^2r_{20}^2}\nu _k(r_{20},y^{})+$$ $$\frac{g^2N_c}{8\pi ^3}_0^y𝑑y^{}e^{2\omega (r_{10})(yy^{})}d^2r_2\frac{r_{10}^2}{r_{12}^2r_{20}^2}\underset{j=1}{\overset{k1}{}}\nu _j(r_{12},y^{})\nu _{kj}(r_{20},y^{}).$$ (65) The total forward scattering amplitude on the nucleus will be given by the expression (54) in which the single dipole interaction $`d^2r_1n_1(r,r_1,y)\tau (r_1)`$ is substituted by the sum of all multidipole interactions $`_k\nu _k(r)`$. Presenting the amplitude in the form (10) we have $$\mathrm{\Phi }(r,y)=\underset{k=1}{}\nu _k(r,y).$$ (66) Summing Eqs. (64) and (65) over $`k`$ we obtain an equation for $`\mathrm{\Phi }`$: $$\mathrm{\Phi }(r_{10},y)=AT(b)\tau (r_{10})e^{2y\omega (r_{10})}+\frac{g^2N_c}{4\pi ^3}_0^y𝑑y^{}e^{2\omega (r_{10})(yy^{})}d^2r_2\frac{r_{10}^2}{r_{12}^2r_{20}^2}\mathrm{\Phi }(r_{20},y^{})$$ $$\frac{g^2N_c}{8\pi ^3}_0^y𝑑y^{}e^{2\omega (r_{10})(yy^{})}d^2r_2\frac{r_{10}^2}{r_{12}^2r_{20}^2}\mathrm{\Phi }(r_{12},y^{})\mathrm{\Phi }(r_{20},y^{}).$$ (67) Comparing this equation with the one derived in , apart from a different inhomogeneous term, which was discussed earlier, we find difference in the spatial arguments of the function $`\mathrm{\Phi }`$. Our $`\mathrm{\Phi }`$ depends only on one such argument: the dipole dimension $`r_{12}`$. The equivalent Kovhegov’s function $`N`$ depends on two spatial arguments: it depends not only on the dipole dimension but also on its center-of-mass coordinate $`b_0`$, not to be confused with the impact parameter $`b`$ which does not enter his equation at all. The dependence on $`b_0`$ should be governed by the inhomogeneous term, which seems to be independent of $`b_0`$ (Eq. (6a) of ). Then the dependence of $`N`$ on the 2nd argument seems to disappear and Kovchegov’s equation coincides with (67). However this contradicts his initial formula (Eq. (4) of ) in which one integrates over all $`b_0`$. Modulo all these (small) inconsistencies and a different inhomogeneous term, our Eq. (67) coincides with Kovchegov’s. ## 8 References 1. A.Mueller, Nucl. Phys.,B415 (1994) 373. 2. A.Mueller and B.Patel, Nucl. Phys.,B425 (1994) 471. 3. M.A.Braun and G.P.Vacca, Eur. Phys. J C6 (1999) 147. 4. R.Peschanski, Phys. Lett. B409 (1997) 491. 5. A.Schwimmer, Nucl. Phys. B94 (1975)445. 6. L.V.Gribov, E.M.Levin an M.G.Ryskin, Nucl. Phys. 188 (1981) 555; Phys. Rep. 100 (1983) 1. 7. J.Bartels and M.Wuesthoff, Z.Phys., C66 (1995) 157. 8. M.A.Braun, Eur. Phys. J C6 (1999) 321. 9. I.Balitsky, hep-ph/9706411; Nucl. Phys. B463 (1996) 99. 10. Yu. Kovchegov, Phys. Rev D60 (1999) 034008. 11 Yu. Kovchegov, preprint CERN-TH/99-166 (hep-ph/9905214). 12.E.Levin and K.Tuchin, preprint DESY 99-108, TAUP 2592-99 (hep-ph/9908317). 13. L.N.Lipatov in: ”Perturbative QCD”, Ed. A.H.Mueller, World Sci., Singapore (1989) 411. 14. M.A.Braun, Eur. Phys. J C6 (1999) 343. 15. N.N.Nikolaev and B.V.Zakharov, Z.Phys. C64 (1994) 631. 16. A.Mueller hep-ph/9904404; 9906322; 9902302. ## 9 Figure captions Fig. 1. The equation summing fan diagrams. Lines represent pomerons. Fig. 2. $`\varphi `$ as a function of momentum at different $`y`$ (in units $`\pi /\alpha _sN`$) for central collisions on lead ($`B=0.12`$, Eq.(359)). Fig. 3. $`\varphi `$ as a function of momentum at different $`y`$ (in units $`\pi /\alpha _sN`$) for peripheral collisions on lead ($`B=0.02`$, Eq.(35)). Fig. 4. Cross-section $`\sigma _L`$ (with $`e^21`$) as a function of $`y`$ (in units $`\pi /\alpha _sN`$) for different $`Q^2`$. Fig. 5. Cross-section $`\sigma _L`$ (with $`e^21`$) as a function of $`Q^2`$ for different $`y`$ (in units $`\pi /\alpha _sN`$). Fig. 6. The structure function of lead $`F_{2A}(x,Q^2)`$ as a function of $`x`$ for different $`Q^2`$. Fig. 7. The structure function of lead $`F_{2A}(x,Q^2)`$ as a function of $`Q^2`$ for different $`x`$. Fig. 8. The gluon density (in units $`\pi ^2N/2\alpha _s`$) as a function of momentum at different rapidities $`\stackrel{~}{y}`$ for central collisions on lead ($`B=0.12`$, Eq. (35)). Fig. 9. The gluon density (in units $`\pi ^2N/2\alpha _s`$) as a function of momentum at different rapidities $`\stackrel{~}{y}`$ for peripheral collisions on lead ($`B=0.02`$, Eq. (35)).
warning/0001/hep-th0001204.html
ar5iv
text
# Untitled Document NSF-ITP-99-146 OHSTPY-HEP-T-99-029 hep-th/0001204 Green-Schwarz String in $`AdS_5\times S^5`$: Semiclassical Partition Function Nadav Drukker,<sup>1,2</sup> David J. Gross,<sup>1</sup> and Arkady A. Tseytlin<sup>3</sup> Also at Lebedev Physics Institute, Moscow and Imperial College, London. <sup>1</sup> Institute for Theoretical Physics University of California Santa Barbara CA, 93106-4030 <sup>2</sup> Department of Physics Princeton University Princeton NJ, 08544 <sup>3</sup> Department of Physics The Ohio State University Columbus OH, 43210-1106 email: drukker, gross@itp.ucsb.edu, tseytlin@mps.ohio-state.edu Abstract A systematic approach to the study of semiclassical fluctuations of strings in $`AdS_5\times S^5`$ based on the Green-Schwarz formalism is developed. We show that the string partition function is well defined and finite. Issues related to different gauge choices are clarified. We consider explicitly several cases of classical string solutions with the world surface ending on a line, on a circle or on two lines on the boundary of $`AdS`$. The first example is a BPS object and the partition function is one. In the third example the determinants we derive should give the first corrections to the Wilson loop expectation value in the strong coupling expansion of the $`𝒩=4`$ SYM theory at large $`N`$. January 2000 1. Introduction The duality between $`AdS_5\times S^5`$ and $`𝒩=4`$ super Yang-Mills theory in four dimensions is the best studied example of the $`AdS`$/CFT correspondence \[1,,2,,3\]. This duality allows the calculation of gauge theory observables at large $`N`$ and large ’t Hooft coupling from perturbative supergravity or string theory. In particular, Wilson loops are described by classical strings that end at the boundary of $`AdS`$ . To extend this duality beyond the supergravity limit it is necessary to learn how to handle strings on this space. Because the background includes a Ramond-Ramond 5-form flux, it is difficult to use the RNS formalism to quantize strings in this geometry. Therefore, one is led to use the Green-Schwarz (GS) action. The first step in this direction was the construction of the classical GS action for strings on this background . Even in flat space the GS action is hard to quantize, except in the light cone gauge. However, this action is perfectly applicable to the perturbative analysis of quantum corrections around a non-trivial “long string” classical solution (assuming the classical bosonic background makes the fermionic kinetic term well-defined). This strategy can be applied in either flat or curved space, and, in particular, is well suited for strings in $`AdS_5\times S^5`$ where there is a natural static solution \[4\] to expand about. The string 2-d loop expansion in $`AdS_5\times S^5`$ is an expansion in powers of $`\alpha ^{}/R^2=\lambda ^{1/2}`$, where $`R`$ is the radius parameter of $`AdS_5\times S^5`$ and $`\lambda `$ is the ’t Hooft coupling: the leading term coming from the classical action is proportional to $`\sqrt{\lambda }`$, the 1-loop correction is just a number, the 2-loop correction will be multiplied by $`\alpha ^{}/R^2=\lambda ^{1/2}`$, etc. The main goal of this paper is to develop technical tools necessary to do calculations of quantum string corrections in $`AdS_5\times S^5`$, at least in the one-loop approximation. This is an important step in the extension of the AdS/CFT correspondence beyond the classical level. Our main motivation is to find the quantum string correction to the Wilson loop expectation value, in particular, the first sub-leading (i.e. $`\lambda `$-independent) correction to the quark anti-quark potential. This problem was first addressed in , where the relevant fermionic operator coming from GS action was derived. An important next step was made in , where the partition function was expressed in terms of operators defined with respect to the induced 2-d geometry. Refs. \[8,,9,,10\] also discussed corrections to the quark anti-quark potential in $`AdS_5\times S^5`$ and in other related geometries. However, all these previous attempts were incomplete as they encountered problems with divergences, gauge fixing, and other subtleties. Our aim is to clarify some of these issues and to set up a consistent framework for performing the semiclassical calculations for the GS string in a curved target space. In particular, we shall explain how the divergence proportional to the world sheet curvature $`R^{(2)}`$ found in is canceled (the cancellation of this divergence in the one-loop approximation in curved target space is essentially the same as in flat space). We will also explain the close relation between the fermionic operators in and in (they correspond to two choices of $`\kappa `$-symmetry gauge). The paper is organized as follows. We start with some general comments about the Green-Schwarz action in flat space, and, in particular, how to use it to calculate quantum corrections to a classical solution. This involves gauge fixing, and determining the measure in the path integral. We will find it most reliable to use conformal gauge, where the path integral measure is best understood \[11,,12\]. Since the theory is critical, the conformal anomalies and, therefore, the 2-d divergences, cancel out. The same mechanism is responsible for the cancellation of leading-order (one-loop) divergences (that are proportional to $`R^{(2)}`$) in curved space as well. In Section 3 we turn to strings on $`AdS_5\times S^5`$. We review the corresponding Brink-DiVecchia-Howe-Polyakov type GS action and explain how to evaluate the quadratic fluctuations around a classical solution. We also comment on the approach based on the Nambu-Goto type action in the static gauge. With careful account of ghosts (and path integral measures) the two approaches should give the same results. We show that as in \[13,,14,,15,,16,,17\] a local Lorentz rotation of GS spinors allows one to systematically transform the quadratic fermionic term in the GS action into the action for a set of 2-d fermions. The problem of computing the partition function is then reduced to the evaluation of determinants of some bosonic and fermionic operators on a 2-d world sheet with an induced metric that is asymptotic to $`AdS_2`$. We study three special examples. In Section 4 we consider a string world surface that ends on a single straight line at the boundary of $`AdS_5`$. The induced metric on the world surface is that of $`AdS_2`$ and the quantum fluctuation fields fit nicely into supersymmetry multiplets on that space. We compute the corresponding vacuum energy and show that it vanishes using a $`\zeta `$-function regularization. The vacuum energy is related to the partition function by a conformal anomaly. Using that we show that the partition function is equal to one. Another case which leads again to $`AdS_2`$ for the induced 2-d geometry is a circular Wilson loop, which we study in Section 5. We comment on the difference between the circular and the straight line cases. In Section 6 we turn to the case of most interest, the surface corresponding to the quark – anti-quark system. Here the induced geometry is more complicated but is still asymptotically $`AdS_2`$. We derive the general expression for the partition function (demonstrating in the process the equivalence of the two $`\kappa `$-symmetry gauges $`\theta ^1=\theta ^2`$ and $`\theta ^1=i\mathrm{\Gamma }_4\theta ^2`$) and discuss evaluation of the numerical coefficient in the corresponding one-loop correction to the $`1/L`$ potential using a crude approximation to the geometry. We summarize our results in Section 7. Some general remarks and explicit calculations are given in Appendices. In Appendix A we review how a determinant of a Laplace operator changes under rescaling of the measure of the fields. The resulting general relations are useful in computing various contributions to the partition function. In Appendix B we present two different calculations of the partition function in the case of $`AdS_2`$ as the induced geometry. Some comments about the fermionic 2-d determinants related to GS action are given in Appendix C. In Appendix D we point out that the expression for the superstring partition function in $`AdS_3\times S^3`$ with RR 2-form background is very similar to the one in the $`AdS_5\times S^5`$ case. Below we shall use the following notation: $`i,j,\mathrm{}=0,1`$ and $`\alpha ,\beta ,\mathrm{}=0,1`$ will denote 2-d world and tangent space indices; $`a,b,\mathrm{}=0,\mathrm{},4`$ and $`p,q,\mathrm{}=1,\mathrm{},5`$ will be the tangent space indices of $`AdS_5`$ and $`S^5`$; $`\widehat{a}=0,1,\mathrm{},9`$ will be the tangent space indices of the 10-d space-time. 2. Green-Schwarz action in flat space Before plunging into discussion of strings in curved target space, it is useful to clarify several general points about the GS action. The flat space GS action of type IIB theory is $$\begin{array}{cc}\hfill S_{\mathrm{flat}}=\frac{1}{2\pi \alpha ^{}}d^2\sigma [& \frac{1}{2}\sqrt{g}g^{ij}\eta _{\widehat{a}\widehat{b}}\left(_ix^{\widehat{a}}i\overline{\theta }^I\mathrm{\Gamma }^{\widehat{a}}_i\theta ^I\right)\left(_jx^{\widehat{b}}i\overline{\theta }^J\mathrm{\Gamma }^{\widehat{b}}_j\theta ^J\right)\hfill \\ & iϵ^{ij}s^{IJ}\overline{\theta }^I\mathrm{\Gamma }_{\widehat{a}}_j\theta ^J(_ix^{\widehat{a}}\frac{1}{2}i\overline{\theta }^K\mathrm{\Gamma }^{\widehat{a}}_i\theta ^K)],\hfill \end{array}$$ where $`\widehat{a}=0,1,\mathrm{},9`$, $`s^{IJ}`$ is defined by $`s^{11}=s^{22}=1`$, $`s^{12}=s^{21}=0`$, $`g_{ij}`$ ($`i,j=0,1`$) is a world-sheet metric with signature $`(+)`$, $`g=detg_{ij}`$, and $`\theta ^I`$ are two left 10-d Majorana Weyl spinors. This action can be considered in either the Polyakov form, with independent 2-d metric (which can be quantized in the conformal gauge) or in the Nambu form, with the induced metric (which can be quantized in the static gauge). When doing semiclassical expansion near a “long string” configuration it may seem natural to use the Nambu formulation, choosing a static gauge. However, the meaning of conformal invariance conditions and the definition of the path integral measure are clear only in the Polyakov formulation. In that case, the 2-d metric, which at the classical level is proportional to the induced metric, should be treated as independent of the coordinates in checking the conformal invariance constraints on the background target space fields (in particular, in proving that conformal anomalies cancel in flat $`D=10`$ space). In the leading 1-loop approximation the Polyakov and Nambu formulations are expected to produce equivalent expressions for the partition function. However, the precise way the divergences cancel may become rather obscure once one sets the metric to be equal to the induced metric, since $`R^{(2)}`$ and $`xx`$ divergences may get mixed up. In particular, the $`R^{(2)}`$ divergences become equivalent to total derivative contributions to $`x`$-dependent divergences and reduce to boundary terms (which may eventually cancel against boundary counterterms). Another important point concerns the distinction between the fermionic kinetic term for GS fermions and for standard 2-d Dirac fermions. As was observed in \[15,,17\] (see also \[19,,13,,16,,14,,20\]) in the case of a flat target space, one may perform a local target space rotation that transforms the quadratic GS fermion term into the 2-d fermion kinetic term. The resulting Jacobian (see, in particular, ) depends on the 2-d metric and its contribution explains why the conformal anomaly of a GS fermion is 4 times bigger than that of a 2-d fermion (which is crucial for understanding how conformal anomalies cancel in $`D=10`$ GS string). Similar remarks apply in the case of curved target spaces. As we shall explicitly discuss below, in some simple cases (like the straight string in $`AdS_5\times S^5`$) the quadratic part of the GS action has already the 2-d fermion form with respect to the curved geometry of induced metric. In other cases one must perform a rotation to express the action in the 2-d fermion form. In the Polyakov formulation with independent 2-d metric, the Jacobian of this must be taken into account for consistent cancellation of conformal anomalies. The contribution of this Jacobian may be non-trivial also in the Nambu formulation where it may depend on the $`x`$-background. 2.1. Quadratic fluctuations near a classical solution The Green-Schwarz action (2.1) is not quadratic in fermions, and is difficult to quantize. One standard way to proceed is to choose a light cone gauge. Alternatively, one may resort to a perturbative expansion in powers of $`\alpha ^{}`$ near a particular classical solution. Since the latter strategy is the only one available in the curved $`AdS_5\times S^5`$ case, we shall employ it below. We concentrate on the one-loop approximation, i.e. on the leading quantum correction to the partition function of the GS string action expanded near a classical solution. With a suitable choice of coordinates, we can write the “long string” classical solution as $$x^0=\sigma ^0,x^1=\sigma ^1.$$ The bosonic part of the action (2.1) is simply the Polyakov action, and it can be quantized in the conformal gauge $`\sqrt{g}g^{ij}=\delta ^{ij}`$. This results in 10 massless world-sheet scalars and two ghosts. Alternatively, one could start with the Nambu form of the action (i.e. first solve for $`g_{ij}`$ and then quantize the theory). In that case we can again expand near (2.1) and choose the static gauge, i.e. eliminate the fluctuations in $`(0,1)`$ directions. Then we are left with just eight transverse scalars. The quadratic term in the fermionic part of the GS action is $$S_{2\mathrm{F}}=\frac{1}{2\pi \alpha ^{}}d^2\sigma L_{2\mathrm{F}}=\frac{i}{2\pi \alpha ^{}}d^2\sigma \left(\sqrt{g}g^{ij}\delta ^{IJ}ϵ^{ij}s^{IJ}\right)\overline{\theta }^I\rho _i_j\theta ^J,$$ where $`g_{ij}`$ can be set equal to $`\eta _{ij}`$ in the conformal gauge. $`\rho _i`$ is the projection of the 10-d Dirac matrices on the world sheet $$\rho _i\mathrm{\Gamma }_{\widehat{a}}_ix^{\widehat{a}}=\mathrm{\Gamma }_i,$$ where the last equality is true for the classical solution (2.1). Then $$L_{2\mathrm{F}}=i\overline{\theta }^1\mathrm{\Gamma }^+_+\theta ^1+i\overline{\theta }^2\mathrm{\Gamma }^{}_{}\theta ^2.$$ This fermionic action is obviously invariant under $`\delta \theta ^1=\mathrm{\Gamma }^+\kappa ^1`$, $`\delta \theta ^2=\mathrm{\Gamma }^{}\kappa ^2`$ which is just the leading-order term in the $`\kappa `$-symmetry transformation rules $$\delta _\kappa \theta ^I=\rho _i\kappa ^{iI}+\mathrm{},\frac{1}{\sqrt{g}}ϵ^{ij}\kappa _j^1=\kappa ^{i1},\frac{1}{\sqrt{g}}ϵ^{ij}\kappa _j^2=\kappa ^{i2}.$$ Since we can represent the (left subspace, 16-component) 10-d Dirac matrices in the form $`\mathrm{\Gamma }_i=\tau _i\times I_8`$, where $`\tau _i`$ are $`2\times 2`$ Dirac matrices and $`I_8`$ is the $`8\times 8`$ unit matrix, the meaning of the $`\kappa `$-symmetry transformations in the present case is simply that $`\theta ^1`$ corresponds to eight left 2-d spinors and $`\theta ^2`$ to eight right 2-d spinors. A natural way to fix $`\kappa `$-symmetry is to set<sup>1</sup> This gauge (considered also in ) is possible only in type IIB theory where the two spinors have the same chirality. This gauge is also natural in connection with the open string theory—in type I theory $`\theta ^1=\theta ^2`$ at the boundary of world sheet. $$\theta ^1=\theta ^2\theta .$$ The remaining degrees of freedom are then 8 real 2-d spinors represented by 16-component left 10-d MW spinor $`\theta `$. The global part of the local $`\kappa `$-symmetry transformation of $`x^\mu `$, that is preserved by the above gauge choice, may then be interpreted as the effective 2-d supersymmetry of the resulting quadratic action, $`\delta x^k=\overline{\theta }^I\mathrm{\Gamma }^k\delta _\kappa \theta ^I`$. This is similar to what happens in the light cone gauge. As we shall see, this simple picture has a direct counterpart in curved case. 2.2. Conformal invariance of GS string in flat space The proof that the fermionic RNS string is conformally invariant at the quantum level is based on adding together the central charges of all the fields: 1 for each scalar boson, $`1/2`$ for each Majorana 2-d fermion, $`26`$ for the conformal ghosts and $`11`$ for the superconformal ghosts. Since conformal anomalies are associated with UV divergences, it is not surprising that the same counting is responsible for the cancellation of logarithmic divergences in the properly defined string partition function on a 2-d surface with any number of holes and handles . This is obvious for the scalar and fermion determinants. For the ghosts, the essential extra ingredient is the need to take into account some global factors in the gauge group measure associated with conformal Killing vectors and/or Teichmüller moduli. Then the logarithmic divergences are again proportional to the total central charge times the Euler number of the Riemann surface and cancel out if $`D=10`$ (or $`D=26`$ in the bosonic string case). The counting for the GS string is different. We describe here only the one-loop approximation. To discuss the cancellation of conformal anomaly in GS string we need to keep the dependence on a generic fiducial metric $`g_{ij}`$ in the action (2.1) and in the norms of the fields ($`d^2\sigma \sqrt{g}\overline{\theta }\theta `$, etc.). Naively, after gauge fixing one gets 10 scalars, 8 Majorana 2-d fermions and the bosonic conformal ghosts. The naive counting would give $`10\times 1+8\times \frac{1}{2}26=12`$. But, in fact, the GS fermionic action depends on the 2-d metric, not as in the case of the standard action for a 2-d spinor (i.e. not through $`\sqrt{g}e_\alpha ^i`$ where $`e_i^\alpha `$ is a zweibein), but rather as a 2-d scalar action (i.e. through $`\sqrt{g}g^{ij}`$). In the conformal gauge $`g_{ij}=e^{2\rho }\eta _{ij},e_i^\alpha =e^\rho \delta _i^\alpha `$ that effectively results in the replacement of $`\rho `$ by $`2\rho `$ in the conformal anomaly term $`(\rho \overline{}\rho )`$ for a 2-d spinor, giving four times bigger a result . Hence the contribution of each of 8 species of GS fermions to the divergence is effectively as of 4 2-d spinors.<sup>2</sup> In more detail, the action for a 2-d spinor is $`d^2\sigma \sqrt{g}g^{ij}\overline{\psi }e_i^\alpha \tau _\alpha _j\psi `$, while for the fermions in the GS action (2.1) the world sheet combination $`\tau _i=e_i^\alpha \tau _\alpha `$ is replaced by the target space one, $`\rho _i=_ix^{\widehat{a}}\mathrm{\Gamma }_{\widehat{a}}`$. In certain cases the two might be equal, but they do not behave the same way under the conformal transformations of the world-sheet metric. The GS fermions $`\theta `$ are world sheet scalars, so their natural measure is $`\theta ^2=d^2\sigma \sqrt{g}\overline{\theta }\theta `$. In the conformal gauge ($`g_{ij}=\sqrt{g}\delta _{ij}`$) the GS fermionic action is $`d^2\sigma \overline{\theta }\rho _\alpha _\alpha \theta `$. Because of the normalization of the $`\theta `$’s, after squaring the fermionic operator we get $`\frac{1}{\sqrt{g}}\frac{1}{\sqrt{g}}\overline{}`$. In the case of the 2-d spinors in the conformal gauge the zweibein contributes to the scaling of the action $`d^2\sigma (\sqrt{g})^{1/2}\overline{\psi }\tau _\alpha _\alpha \psi `$, $`\psi ^2=d^2\sigma \sqrt{g}\overline{\psi }\psi `$. If we rescale $`\psi `$ to make the action $`g`$-independent as in $`\theta `$ case we get $`d^2\sigma \overline{\psi }\tau _\alpha _\alpha \psi `$, $`\psi ^2=d^2\sigma (\sqrt{g})^{1/2}\overline{\psi }\psi `$. The difference compared to $`\theta `$ is now only in the norm. The corresponding 2-nd order operator is $`\frac{1}{(\sqrt{g})^{1/2}}\frac{1}{(\sqrt{g})^{1/2}}\overline{}`$. It is the difference in the measure factor that now leads to different anomalies: for the operator $`ff\overline{}`$ the conformal anomaly in the partition function is $`\mathrm{exp}(\frac{1}{48\pi }d^2\sigma |\mathrm{ln}f|^2)`$, so that the difference between the “scalar” and “2-d spinor” descriptions produces indeed the factor of 4 in the conformal anomaly. Then the count of anomalies in the GS string goes as follows $$1026+8\times 4\times \frac{1}{2}=0.$$ Essentially equivalent arguments (based on separating the metric dependence in a WZ type Jacobian contribution due to a rotation of spinors) which explains why conformal anomaly cancels in $`D=10`$ in GS string were given in \[15,,17,,14\] (see also Appendix C). Since in a covariant regularization the cutoff is coupled to the conformal factor, the cancellation of conformal anomalies should imply also the cancellation of the UV divergences. 2.3. Cancellation of quantum correction to straight string configuration A natural classical “long string” solution in flat space is (2.1), or restoring the dimensional parameters, $$x^0=T\tau ,x^1=L\sigma ,\sigma (\frac{1}{2},\frac{1}{2}).$$ The fluctuations of the $`d2=8`$ transverse bosonic coordinates (which are periodic in $`\sigma ^0`$ and Dirichlet in $`\sigma ^1`$ directions) give for $`T\mathrm{}`$ \[23\] $$W=\mathrm{ln}Z=(d2)W_0,W_0=\frac{1}{2}[\mathrm{log}det(^2)]_T\mathrm{}=\frac{\pi }{24}\frac{T}{L}.$$ In the flat-space superstring case this contribution is canceled by the contribution of the fermionic determinant: the total effective number of transverse world-sheet degrees of freedom is equal to zero (as in any flat-space string theory without tachyons ) because of the effective 2-d supersymmetry present after choosing $`\theta ^1=\theta ^2`$ and expanding to quadratic order near (2.1). Indeed, the induced metric and zweibein are flat, and thus, apart from the subtlety with cancellation of conformal anomalies discussed above, the GS fermionic determinant is the same as for eight 2-d spinors. 3. Quadratic fluctuations of superstring in $`AdS_5\times S^5`$ We now turn to the discussion of the one-loop approximation to the partition function of GS superstring in $`AdS_5\times S^5`$. We start with the Polyakov form of GS action in conformal gauge, expand near a general classical solution, and explicitly check conformal invariance to 1-loop order. We shall also comment on the result obtained by starting with the Nambu-type action and using static gauge. In the following sections we give examples of particular symmetric solutions. In the context of the $`AdS`$/CFT duality, expansion about classical solutions of the string action, namely minimal surfaces, corresponds to computing expectation values of Wilson loop operators in the dual gauge theory . The expectation value of the Wilson loop is given by $$W=[dx][d\theta ][dg]e^S,$$ where $`S`$ is the string action in $`AdS_5\times S^5`$ and the path integral is over all embeddings of the string into $`AdS_5\times S^5`$ with proper boundary conditions (the string world surface should end along the loop at the boundary of $`AdS_5`$ \[4,,25\]). Here we assumed the Polyakov form, where, in general, one is to integrate over the moduli of 2-d metrics. 3.1. The action The bosonic part of the action for a string in $`AdS_5\times S^5`$ is<sup>3</sup> For a string representing a Wilson loop of SYM theory and ending at the boundary the classical action is actually a particular Legendre transform of the area , but that does not affect the discussion of quantum fluctuations. $$S_\mathrm{B}=\frac{R^2}{4\pi \alpha ^{}}d^2\sigma \sqrt{g}g^{ij}G_{\mu \nu }(x)_ix^\mu _jx^\nu .$$ We have removed the dependence on the $`AdS`$ scale $`R`$ ($`R^4=4\pi \alpha ^2g_sN`$) from the space-time metric $`G_{\mu \nu }`$ ($`m=1,\mathrm{},4`$) $$ds^2=\frac{1}{w^2}\left(dw^2+dx^mdx^m\right)+d\mathrm{\Omega }_5^2.$$ The leading behavior at large ’t Hooft coupling $`\lambda `$ is the exponent of the classical action, which is proportional to $`\frac{R^2}{\alpha ^{}}=\sqrt{\lambda }`$. The string expansion is in inverse powers of $`\sqrt{\lambda }`$. In most of the paper we set $`R=1`$, but it is easy to restore the dependence on $`R`$ when necessary. The structure of the full covariant GS string action in $`AdS_5\times S^5`$ is rather complicated , but the part quadratic in $`\theta ^I`$ is simple and is a direct generalization of the quadratic term in the flat-space GS action (2.1) $$S_{2\mathrm{F}}=\frac{i}{2\pi \alpha ^{}}d^2\sigma (\sqrt{g}g^{ij}\delta ^{IJ}ϵ^{ij}s^{IJ})\overline{\theta }^I\rho _iD_j\theta ^J.$$ Here $`\rho _i`$ are again projections of the 10-d Dirac matrices, $$\rho _i\mathrm{\Gamma }_{\widehat{a}}E_\mu ^{\widehat{a}}_ix^\mu =(\mathrm{\Gamma }_aE_\mu ^a+\mathrm{\Gamma }_pE_\mu ^p)_ix^\mu ,$$ and $`E_\mu ^{\widehat{a}}`$ is the vielbein of the 10-d target space metric, $`G_{\mu \nu }=E_\mu ^{\widehat{a}}E_\nu ^{\widehat{b}}\eta _{\widehat{a}\widehat{b}}`$. The covariant derivative $`D_i`$ is the projection of the 10-d derivative $`D_\mu =_\mu +\frac{1}{4}\mathrm{\Omega }_\mu ^{\widehat{a}\widehat{b}}\mathrm{\Gamma }_{\widehat{a}\widehat{b}}\frac{1}{85!}\mathrm{\Gamma }^{\mu _1\mathrm{}\mu _5}\mathrm{\Gamma }_\mu e^\varphi F_{\mu _1\mathrm{}\mu _5}`$ ($`\mathrm{\Omega }_\mu ^{\widehat{a}\widehat{b}}`$ is the spin connection and $`F_{\mu _1\mathrm{}\mu _5}`$ the RR 5-form potential) which appears, e.g., in the Killing spinor equation of type IIB supergravity. It has the following explicit form <sup>4</sup> The 10-d Dirac matrices are split in the ‘5+5’ way, $`\mathrm{\Gamma }^a=\gamma ^a\times I_4\times \sigma _1`$, $`\mathrm{\Gamma }^p=I_4\times \gamma ^p\times \sigma _2,`$ where $`\sigma _k`$ are Pauli matrices and $`\gamma ^a`$, $`\gamma ^p`$ are $`4\times 4`$ matrices (corresponding to tangent spaces of $`AdS_5`$ and $`S^5`$ factors) and $`I_4`$ is $`4\times 4`$ unit matrix (see for details on notation; we use index $`p=1,\mathrm{},5`$ instead of $`a^{}`$ in ). A $`D=10`$ positive chirality 32-component spinor $`\mathrm{\Psi }`$ is decomposed as follows: $`\mathrm{\Psi }=\psi \times \psi ^{}\times \left(\genfrac{}{}{0pt}{}{1}{0}\right)`$. In equations written in the 32-component spinor form, $`\theta ^I`$ stands for two positive chirality spinors $`\left(\genfrac{}{}{0pt}{}{\theta ^I}{0}\right)`$, where $`\theta ^I`$ are 16-component spinors used below. The Majorana condition $`\overline{\mathrm{\Psi }}\mathrm{\Psi }^{}\mathrm{\Gamma }^0=\mathrm{\Psi }^T𝒞`$, $`𝒞C\times C^{}\times i\sigma _2`$, then takes the form $`\overline{\theta }_{\alpha \alpha ^{}I}(\theta ^{I\beta \beta ^{}})^{}(\gamma ^0)_\alpha ^\beta \delta _\alpha ^{}^\beta ^{}=\theta ^{I\beta \beta ^{}}C_{\beta \alpha }C_{\beta ^{}\alpha ^{}}^{}`$. Here $`C`$ and $`C^{}`$ are the charge conjugation matrices of $`so(4,1)`$ and $`so(5)`$ used to raise and lower spinor indices. Note that $`C\times C^{}`$ is symmetric, i.e. $`\overline{\theta }\theta =0`$. In the expressions below $`\theta ^I`$ may be thought of as 5-d spinors with an extra ‘spectator’ 5-d spinor index, and we shall assume that $`\gamma ^a`$ and $`\gamma ^p`$ stand for $`\gamma ^a\times I_4`$ and $`I_4\times \gamma ^p`$. $`C\gamma ^{a_1\mathrm{}a_n}`$ are symmetric (antisymmetric) for $`n=2,3`$ mod 4 ($`n=0,1`$ mod 4). The same properties are valid for $`C^{}\gamma ^{p_1\mathrm{}p_n}`$. We also assume that $`\gamma _0^{}=\gamma _0`$, $`\gamma _m^{}=\gamma _m`$ $`(m=1,2,3,4)`$. $$D_i\theta ^I\left(\delta ^{IJ}𝒟_i\frac{1}{2}iϵ^{IJ}\stackrel{~}{\rho }_i\right)\theta ^J,𝒟_i=_i+\frac{1}{4}_ix^\mu \mathrm{\Omega }_\mu ^{\widehat{a}\widehat{b}}\mathrm{\Gamma }_{\widehat{a}\widehat{b}},$$ where the $`\stackrel{~}{\rho }_i`$ term originates from the coupling to the RR field strength, $$\stackrel{~}{\rho }_i\left(\mathrm{\Gamma }_aE_\mu ^a+i\mathrm{\Gamma }_pE_\mu ^p\right)_ix^\mu .$$ Note that $`\stackrel{~}{\rho }_i`$ is not identical to $`\rho _i`$, unless one is expanding near a classical solution that is constant on $`S^5`$. In general, there is a factor of $`R^1`$ in front of the ‘mass term’, so that this term disappears in the flat-space limit. 3.2. Expanding about a classical solution We first consider the bosonic sector. Expand the Polyakov action (3.1) about a classical solution $$x^\mu \overline{x}^\mu +\xi ^\mu ,g_{ij}g_{ij}+\chi _{ij},$$ $$g_{ij}=e^{2\lambda }h_{ij},h_{ij}G_{\mu \nu }(\overline{x})_i\overline{x}^\mu _j\overline{x}^\nu .$$ The classical value of the metric may, in general, differ from the induced metric $`h_{ij}`$ by an arbitrary conformal factor $`\lambda `$. We fix the 2-d diffeomorphism invariance by imposing the conformal gauge conditions on the fluctuations of the metric $$\chi _{ij}=\kappa g_{ij},\mathrm{i}.\mathrm{e}.g_{ij}(1+\kappa )g_{ij}.$$ The remaining conformal degree of freedom of the metric should decouple as in flat 10-d space because of the conformal invariance of type IIB string theory on $`AdS_5\times S^5`$ background \[5,,26\]. To check this we treat $`g_{ij}`$ as an arbitrary background metric, not identifying it (both in the action and in the path integral measure) with $`h_{ij}`$. Let us introduce the tangent-space components of the fluctuation fields which have the canonical norms $$\begin{array}{cc}\hfill \zeta ^a=E_\mu ^a\xi ^\mu ,\zeta ^p=E_\mu ^p\xi ^\mu ,G_{\mu \nu }=E_\mu ^aE_\nu ^a+E_\mu ^pE_\nu ^p,& \\ \hfill \zeta ^a^2=d^2\sigma \sqrt{g}\zeta ^a\zeta ^a,\zeta ^p^2=d^2\sigma \sqrt{g}\zeta ^p\zeta ^p.& \end{array}$$ $`\zeta ^a`$ and $`\zeta ^q`$ are the fluctuations of the $`AdS_5`$ and $`S^5`$ coordinates respectively (the tangent-space 5-d indices $`a,b=0,\mathrm{},4`$ and $`p,q=1,\mathrm{},5`$ are raised by the flat 5-d metrics). We then get the following action for the quadratic fluctuations (we absorb $`\frac{1}{2\pi \alpha ^{}}`$ by rescaling the quantum fields) $$S_{2\mathrm{B}}=\frac{1}{2}d^2\sigma \sqrt{g}\left(g^{ij}D_i\zeta ^aD_j\zeta ^a+X_{ab}\zeta ^a\zeta ^b+g^{ij}D_i\zeta ^qD_j\zeta ^q+X_{pq}\zeta ^p\zeta ^q\right),$$ $$X_{ab}=g^{ij}\eta _i^c\eta _j^dR_{acbd},X_{pq}=g^{ij}\eta _i^r\eta _j^sR_{prqs}.$$ Here $$\eta _i^a_i\overline{x}^\mu E_\mu ^a,\eta _i^p_i\overline{x}^\mu E_\mu ^p,$$ are the projection of the $`AdS_5`$ and $`S^5`$ vielbeins on the world sheet. $`D_i`$ is the covariant derivative containing the projection of the target space spin connection, $$D_i\zeta ^a=_i\zeta ^a+w_i^{ab}\zeta ^b,w_i^{ab}=_i\overline{x}^\mu \mathrm{\Omega }_\mu ^{ab},$$ where $`\mathrm{\Omega }_\mu ^{ab}`$ is the spin connection of $`AdS_5`$, and similarly for $`S^5`$. For example, the $`AdS_5`$ part of the connection is $`\mathrm{\Omega }_\mu ^{m4}=w^1\delta _\mu ^m`$, where $`4`$ stands for the radial direction and $`m=0,1,2,3`$. In general, there will be two types of divergences – depending on the background $`\overline{x}`$ field $`O(\overline{x}\overline{x})`$ (i.e. renormalization of the target space metric) and proportional to the curvature $`R^{(2)}`$ of the fiducial 2-d metric $`g_{ij}`$ (i.e. renormalization of the dilaton). To check the conformal invariance we need to use the fact that for $`AdS_5\times S^5`$ $$R_{acbd}=\delta _{ab}\delta _{cd}+\delta _{ad}\delta _{cb},R_{prqs}=\delta _{pq}\delta _{rs}\delta _{ps}\delta _{rq}.$$ Then $$X^{ab}=g^{ij}\eta _i^c\eta _j^c\delta ^{ab}g^{ij}\eta _i^a\eta _j^b,X^{pq}=g^{ij}\eta _i^r\eta _j^r\delta ^{pq}+g^{ij}\eta _i^p\eta _j^q.$$ The $`\overline{x}`$-dependent logarithmic UV divergences coming from (3.1) are proportional to $$\mathrm{tr}X=4g^{ij}\left(\eta _i^a\eta _j^a\eta _i^p\eta _j^p\right).$$ This gives the Ricci tensor dependent of the conformal invariance equation, $`R_{\widehat{a}\widehat{b}}\frac{1}{44!}F_{\widehat{a}\mathrm{}.}F_{\widehat{b}\mathrm{}.}=0`$. The 5-form dependent part will come from the fermionic contribution. $`\kappa `$-symmetry transformations in curved $`AdS_5\times S^5`$ space which leave (3.1) invariant have a form similar to flat-space transformations (2.1) $$\delta _\kappa \theta ^I=\stackrel{~}{\rho }_i^{}\kappa ^{iI}+\mathrm{},$$ where $$\frac{1}{\sqrt{g}}ϵ^{ij}\kappa _j^1=\kappa ^{i1},\frac{1}{\sqrt{g}}ϵ^{ij}\kappa _j^2=\kappa ^{i2},$$ and (cf. (3.1)) $$\stackrel{~}{\rho }_i^{}=\left(\mathrm{\Gamma }_aE_\mu ^ai\mathrm{\Gamma }_pE_\mu ^p\right)_ix^\mu .$$ Fixing the $`\kappa `$-symmetry gauge $$\theta ^1=\theta ^2=\theta ,$$ the quadratic part of the fermionic action (3.1)–(3.1) is found to be $$S_{2\mathrm{F}}=2id^2\sigma \left(\sqrt{g}g^{ij}\overline{\theta }\rho _i𝒟_j\theta \frac{i}{2}ϵ^{kj}\overline{\theta }\rho _k\stackrel{~}{\rho }_j\theta \right).$$ We shall first be interested in $`\overline{x}\overline{x}`$ divergences so that it should not be necessary to distinguish between $`g_{ij}`$ and the induced metric, therefore $$\rho _{(i}\rho _{j)}=g_{ij}=G_{\mu \nu }_i\overline{x}^\mu _j\overline{x}^\nu =\eta _i^a\eta _j^a+\eta _i^p\eta _j^p.$$ In general, given the operator $`𝒪=i\rho ^k𝒟_k+M`$, the logarithmic divergence in $`\frac{1}{2}\mathrm{ln}det(𝒪𝒪^{})`$ is proportional to<sup>5</sup> Starting with $`(i\rho ^k_k+M)(i\rho ^n_n+M^{})`$ one is to note that in the general case of $`\rho ^kM^{}+M\rho ^k0`$ one is to introduce an additional connection to put the resulting operator in the standard form $`D^2+X`$. $$\frac{1}{4}g^{ij}\mathrm{tr}(\rho _iM\rho _jM+\rho _iM^{}\rho _jM^{}),$$ where in the present case $`M=\frac{1}{2}\frac{ϵ^{kj}}{\sqrt{g}}\rho _k\stackrel{~}{\rho }_j,M^{}=\frac{1}{2}\frac{ϵ^{kj}}{\sqrt{g}}\stackrel{~}{\rho }_k^{}\rho _j`$, $`\rho _i=\rho _i^{}`$. After some algebra<sup>6</sup> If the background does not depend on $`S^5`$, then $`\rho _i=\stackrel{~}{\rho }_i`$ and the calculation is trivial. one finds that (3.1) reduces to $$4\left[det\left(\eta _i^a\eta _j^a\right)det\left(\eta _i^p\eta _j^p\right)\right]=4g^{ij}\left(\eta _i^a\eta _j^a\eta _i^p\eta _j^p\right).$$ This combination determines the fermionic contribution to the $`\overline{x}`$-dependent logarithmic divergence, and it exactly cancels the bosonic contribution (3.1). As usual, fixing conformal gauge produces bosonic ghosts which are 2-d vectors $$S_{\mathrm{gh}}=\frac{1}{2}d^2\sigma \sqrt{g}g^{ij}\left(g^{kl}_kϵ_i_lϵ_j\frac{1}{2}R^{(2)}ϵ_iϵ_j\right),$$ where $`R^{(2)}`$ is the scalar curvature of the 2-d metric $`g_{ij}`$. The quadratic and linear divergences cancel between bosons and fermions because of the matching of the number of degrees of freedom. The coefficients of the logarithmic divergence for the above system of fields (the Seeley coefficients of the corresponding second order Laplace operators) are $$b_{2\mathrm{B}}=10\times \frac{R^{(2)}}{6}\mathrm{tr}X,b_{2\mathrm{gh}}=2\times \frac{R^{(2)}}{6}R^{(2)},$$ $$b_{2\mathrm{F}}=8\times \frac{R^{(2)}}{3}+\mathrm{tr}X.$$ Here we took into account that since the kinetic part of the fermionic operator depends on the background 2-d metric through $`\sqrt{g}g^{ij}`$, the $`R^{(2)}`$-dependent part of its divergence and conformal anomaly coefficient is four times greater than for a 2-d Majorana fermion, just like in the flat GS string case (2.1) (this difference may be attributed to the contribution of the Jacobian of a local rotation that transforms $`\rho _i`$ into 2-d Dirac matrices contracted with zweibein, see Appendix C). The total divergence coefficient is then $$b_2^{\mathrm{total}}=\left(102+8\times 2\right)\times \frac{R^{(2)}}{6}R^{(2)}=3R^{(2)}.$$ As was mentioned above, at 1-loop order the argument for the cancellation of logarithmic divergences is identical to the argument in the case of the flat GS string, where (3.1) is also valid. Integrating over the scalar curvature on a closed surface will give the Euler character $`d^2\sigma \sqrt{g}\mathrm{\hspace{0.17em}3}R^{(2)}=12\pi \chi `$. The same is true on a surface with boundary where, as was shown in , all the factors of $`R^{(2)}`$ are accompanied by the appropriate boundary term. Now one should remember that the cutoff dependent factors in the conformal Killing vector and/or Teichmüller measure exactly cancels this divergence, so the final result is $`(D10)\chi `$, namely zero. This is, of course, consistent with the cancellation of the total $`\overline{x}`$-independent conformal anomaly, or the central charge, ensuring that the dilaton equation is satisfied. Note that this is just a consequence of working in the critical dimension. Thus, we have confirmed that the theory is conformal at one loop. As was argued in \[5,,26\], this should be true to all orders in $`\alpha ^{}`$ expansion (for example, the first non-trivial correction to the central charge vanishes because the Ricci scalar of the target space metric is zero, $`R_{tot}=R_{AdS_5}+R_{S^5}=0`$, etc.). 3.3. Nambu-type action in the static gauge Alternatively, one may start with the corresponding Nambu-Goto form of the GS action (with no independent 2-d metric). This action is highly non-linear, but in the quadratic approximation it is straightforward to determine the 2-d operators of small fluctuations of a string in curved background. Here it is natural to choose the static gauge to fix the diffeomorphisms, i.e. to identify the world sheet coordinates with the two target space coordinates and demand that there are no fluctuations in those directions. The ghost determinant is then “local”, i.e. is a determinant of an operator of multiplication by a function (which needs a regularization and may still produce non-trivial contribution to partition function). Following the standard procedure (for a careful treatment see, e.g., ), fixing the static gauge $`\delta x^k=0`$, $`k=0,1`$, produces the ghost determinant $$\mathrm{\Delta }_{\mathrm{gh}}^1=[dϵ]\mathrm{exp}\left(\frac{1}{2}\delta _ϵx^k^2\right).$$ The path integral over the 2-d diffeomorphism parameters $`ϵ^i`$ is defined using the norm $$ϵ^2=d^2\sigma \sqrt{h}h_{ij}ϵ^iϵ^j.$$ Here $`h_{ij}`$ is the induced metric (3.1) $$h_{ij}=G_{\mu \nu }(\overline{x})_i\overline{x}^\mu _j\overline{x}^\nu =\eta _i^{\widehat{a}}\eta _j^{\widehat{a}}.$$ Explicit evaluation of (3.1) gives $$\delta _ϵx^k^2=d^2\sigma \sqrt{h}G_{kl}(\overline{x})(ϵ^i_i\overline{x}^k)(ϵ^j_j\overline{x}^l)=d^2\sigma \sqrt{h}h_{ij}^{}ϵ^iϵ^j,$$ $$h_{ij}^{}G_{kl}(\overline{x})_i\overline{x}^k_j\overline{x}^l,$$ so that $$\mathrm{\Delta }_{\mathrm{gh}}=[det(h_{ik}^{}h^{kj})]^{1/2}.$$ As we will see on the examples discussed below, the most natural regularization of this “local” determinant is by changing the normalization of some of the fluctuating fields. Another possible gauge is to remove the vielbein components of the longitudinal fluctuations. It is easy to see, by the same calculation, that the ghost determinant is equal to 1 in that gauge. The resulting bosonic action is a modification of (3.1) $$S=\frac{1}{2}d^2\sigma \sqrt{h}\left(h^{ij}D_i\zeta ^{\overline{a}}D_j\zeta ^{\overline{a}}+\overline{X}_{\overline{a}\overline{b}}\zeta ^{\overline{a}}\zeta ^{\overline{b}}\right),$$ where $`\zeta ^{\overline{a}}`$ are the fields representing the transverse fluctuations. $`\overline{X}`$ is not the same as $`X`$ (in the simple examples discussed below $`\mathrm{tr}\overline{X}=\mathrm{tr}X+R^{(2)}`$). The fermions are treated in the same way as before, so squaring the fermionic operator gives a mass term whose trace is equal to $`\mathrm{tr}X`$. The non-trivial $`O(\overline{x}\overline{x})`$ part of divergences cancels again, while the remaining $`R^{(2)}(h)`$ part (which, in the presence of a boundary should be accompanied by an appropriate boundary term to give the Euler number) should be canceled in $`D=10`$ by appropriate measure factor contributions, as happens in conformal gauge . It should be stressed that, while the result of a semiclassical computation in the Nambu action case should be equivalent to the one in the Polyakov action case , a careful definition of the path integration measure is non-trivial in the Nambu case. For that reason we prefer to use the Polyakov definition of the string partition function which is well-defined. It should be clear that the problem of cancellation of $`R^{(2)}`$ divergences is exactly the same as in the case of the Nambu action in flat space, and thus has nothing to do with peculiarities of the $`AdS_5\times S^5`$ background. This resolves the puzzle of the apparent non-cancellation of logarithmic divergences that was encountered in , which is revealed to be an artifact of the use of the Nambu-type formulation without including the additional measure contributions to the divergences and ignoring the subtleties of the divergences/conformal anomaly cancellation in the flat-space GS superstring theory. We shall see that in the cases of interest the kinetic operator of GS fermions takes (after a local rotation) the form of the 2-d Dirac operator in the curved 2-d geometry defined by the induced metric. In view of the above discussion, the fact that when one directly evaluates the divergences one seems to find that the $`R^{(2)}`$-terms do not cancel is an artifact of not distinguishing between the generic and the induced metrics. These topological divergences are, in any case, irrelevant for the evaluation of the non-trivial part of the partition function which determines the correction to the Wilson loop expectation value, or to the $`1/L`$ potential. The issue of divergences may be avoided altogether, by normalizing, as we suggest below, the partition function to its value for some standard background. Once this issue has been clarified, one should be able to use the static gauge expressions for the small-fluctuation determinants, as they sometimes turn out to be simpler than the analogous expressions in the Polyakov formulation in conformal gauge. In what follows we shall not distinguish between the generic fiducial metric $`g_{ij}`$ and the induced metric $`h_{ij}`$ (using always the notation $`g_{ij}`$ for the 2-d metric). 3.4. Relating quadratic GS fermion term to 2-d Dirac fermion action Before turning to specific examples let us make some general comments on how one can put the quadratic fermionic term (3.1) in the GS action in curved target space background into the standard kinetic term for a set of 2-d fermions defined on a curved 2-d space. The main idea is to apply a local target space Lorentz rotation to GS spinor $`\theta `$, as discussed previously in the case of the heterotic string in flat \[13,,14,,15\] and curved \[16,,17\] spaces. We shall concentrate on the derivative term in the fermionic action (3.1). It should be noted that the presence of the second “mass” term in (3.1) which originates from the coupling to RR background and which is absent in the heterotic case will not allow to compute the resulting 2-d fermion determinant in a closed form using the standard anomaly arguments. Ignoring the distinction between $`g_{ij}`$ and the induced metric (3.1) we can write the derivative term in (3.1) as $$S_{2\mathrm{F}}^{(deriv.)}=2id^2\sigma \sqrt{g}g^{ij}\overline{\theta }\rho _i𝒟_j\theta =id^2\sigma \sqrt{g}g^{ij}(\overline{\theta }\rho _i_j\theta _j\overline{\theta }\rho _i\theta ).$$ Let us introduce the tangent $`t_\alpha ^\mu `$ ($`\mu =0,1,\mathrm{},9`$, $`\alpha =0,1`$) and normal $`n_s^\mu `$ ($`s=1,\mathrm{},8`$) vectors to the world surface which form orthonormal 10-d basis ($`g_{ij}=e_i^\alpha e_j^\beta \eta _{\alpha \beta }`$) $$t_\alpha ^\mu =e_\alpha ^i_i\overline{x}^\mu ,(t_\alpha ,t_\beta )=\eta _{\alpha \beta },(t_\alpha ,n_s)=0,(n_s,n_u)=\delta _{su},$$ where $`(a,b)=G_{\mu \nu }a^\mu b^\nu `$. Then one can make a local $`SO(1,9)`$ rotation of this basis which transforms the set of $`\sigma `$-dependent 10-d Dirac matrices (see (3.1)) into the 10 constant Dirac matrices $$\rho _\alpha (\sigma )=e_\alpha ^i\rho _i=S(\sigma )\mathrm{\Gamma }_\alpha S^1(\sigma ),\rho _s(\sigma )=n_s^\mu E_\mu ^{\widehat{a}}\mathrm{\Gamma }_{\widehat{a}}=S(\sigma )\mathrm{\Gamma }_sS^1(\sigma ).$$ One may further choose a representation in which $`\mathrm{\Gamma }_\alpha =\tau _\alpha \times I_8`$, where $`\tau _\alpha `$ are 2-d Dirac matrices. Depending on the specific embedding and particular curved target space metric, one may then be able to write the action (3.1) as the action for 2-d Dirac fermions coupled to curved induced 2-d metric and interacting with some gauge fields (coming from $`S^1dS`$). Simple examples when this happens will be discussed below. We shall consider embeddings of the string world sheet into the $`AdS_3`$ part of the $`AdS_5`$ space, so there will be only one normal direction and the extra normal bundle 2-d gauge connection will be absent (cf. \[13,,14,,16\]). In this 3-dimensional embedding case with non-chiral 2-d fermions the Jacobian associated with the local Lorentz rotation will be trivial (see also Appendix C). 4. One-loop approximation near the straight string configuration: Supersymmetric field theory on $`AdS_2`$ The simplest classical solution for string in $`AdS_5\times S^5`$ is a straight string with the world surface spanned by the radial direction of $`AdS_5`$ and time. The Euclidean solution and the corresponding induced metric are $$\tau =x^0,\sigma =x^4=w,ds^2=\frac{1}{\sigma ^2}(d\tau ^2+d\sigma ^2).$$ The induced metric on the world sheet is that of $`AdS_2`$, with constant negative curvature $`R^{(2)}=2`$ (the radius of $`AdS_5`$ is $`R=1`$). This solution represents a single straight Wilson line running along the Euclidean time direction. This is a BPS object in string theory, it corresponds to a static fundamental string stretched between a single D3-brane and $`N`$ coinciding D3-branes. Therefore, one would expect that the partition function be equal to 1. The properly defined (subtracted) classical string action evaluated on (4.1) indeed vanishes, and we shall evaluate the 1-loop correction to the partition function. As we shall show, the corresponding 1-loop correction to the vacuum energy defined with respect to a certain time-like Killing vector vanishes. Relating the vacuum energy to the partition function using a conformal rescaling argument (and the fact that the total conformal anomaly is zero) we conclude that $`Z=1`$. It should be mentioned that while the (properly defined) vacuum energy of a supersymmetric field theory in $`AdS`$ space should vanish, this does not automatically imply (in contrast to what happens in flat space) that the partition function of such theory should be equal to 1 (cf. \[30,,31\]). In the present case this happens only with the inclusion of the appropriate ghosts and longitudinal modes. The calculation of the partition function is rather subtle, and depends on a regularization prescription. Let us note also that the point of view of physical applications, the precise value of $`Z`$ (which is simply a constant) is not actually important, and one may normalize with respect to it in computing $`Z`$ for more general string configurations. Apart from being the simplest example, there are other reasons why the analysis of the straight string case is of interest. Any smooth Wilson loop looks in the UV region like a straight line. In the present set-up this translates into the behavior of the minimal surface near the boundary of $`AdS_5`$ space. In the general case one will have to calculate the partition function for a complicated two dimensional field theory. But asymptotically the minimal surface will approach $`AdS_2`$, and the small fluctuation operators (in particular, the asymptotic values of the masses of the fluctuation fields) will also be the same as for a straight string. Many subtleties related to divergences and asymptotic boundary conditions are already present in this example, and they can be automatically avoided in more general cases by normalizing with respect to the partition function of the straight string. 4.1. The action and multiplet structure The bosonic part of the action for small fluctuations in conformal gauge is (3.1) $$S_{2\mathrm{B}}=\frac{1}{2}d^2\sigma \sqrt{g}\left(g^{ij}D_i\zeta ^aD_j\zeta ^a+X_{ab}\zeta ^a\zeta ^b+g^{ij}D_i\zeta ^pD_j\zeta ^p\right),$$ where in the present case $$\eta _0^a=_0\overline{x}^\mu E_\mu ^a=w^1(1,0,0,0,0),\eta _1^a=_1\overline{x}^\mu E_\mu ^a=w^1(0,0,0,0,1).$$ $`\eta _i^a`$ with $`a=0,4`$ is thus just a vielbein of the induced 2-d metric $`g_{ij}`$, $$\eta _i^a=e_i^\alpha ,a=0,4,\alpha =0,1,$$ so that $$X_{ab}=\mathrm{diag}(1,2,2,2,1).$$ The only nonzero connection in $`D_i`$ (3.1) is $`w_0^{04}=w^1`$, i.e. $$\begin{array}{cc}\hfill D_0\zeta ^0=_0\zeta ^0w^1\zeta ^4,D_0\zeta ^4=_0\zeta ^4+w^1\zeta ^0.& \end{array}$$ The natural norms for the fields are $$\zeta ^{\widehat{a}}^2=d^2\sigma \sqrt{g}\zeta ^{\widehat{a}}\zeta ^{\widehat{a}}=𝑑\tau 𝑑\sigma \frac{1}{\sigma ^2}\zeta ^{\widehat{a}}\zeta ^{\widehat{a}}.$$ The ghost action is the same as in (3.1), i.e. $$\frac{1}{2}d^2\sigma \sqrt{g}(^iϵ^\alpha _iϵ^\alpha \frac{1}{2}R^{(2)}ϵ^\alpha ϵ^\alpha ),$$ where we defined $`ϵ^\alpha =e_i^\alpha ϵ^i`$ with flat 2-d tangent space indices, and $`_i`$ includes the world sheet Lorentz connection. Because of the direct embedding of the world sheet into the target space there are some extra simplifications. The projection of the target space connection on the world sheet $`w_i^{ab}`$ is the same as the spin connection of the induced metric appearing in $`_i`$. In addition, $`\frac{1}{2}R^{(2)}=1`$. Therefore, the action of the ghosts is identical to the action of the longitudinal modes $`\zeta ^0`$, $`\zeta ^4`$, but the boundary conditions are different . Before $`\kappa `$-symmetry gauge fixing the fermionic Lagrangian (3.1) is (here we use Minkowski notation) $$L_{2\mathrm{F}}=i\left(\sqrt{g}g^{ij}\delta ^{IJ}ϵ^{ij}s^{IJ}\right)\overline{\theta }^I\rho _iD_j\theta ^J,$$ where in the present case $$\begin{array}{cc}\hfill \rho _i& =e_i^\alpha \rho _\alpha =\eta _i^a\mathrm{\Gamma }_a=\{\begin{array}{ccc}w^1\mathrm{\Gamma }_0& \text{for}& i=0,\\ w^1\mathrm{\Gamma }_4& \text{for}& i=1,\end{array}\hfill \\ \hfill D_i\theta ^J& =\widehat{}_i\theta ^J\frac{i}{2}ϵ^{JK}\rho _i\theta ^K,\hfill \\ \hfill \widehat{}_0& =_0\frac{1}{2w}\mathrm{\Gamma }_{04},\widehat{}_1=_1.\hfill \end{array}$$ We choose again the gauge $`\theta ^1=\theta ^2=\theta `$. Then $$L_{2\mathrm{F}}=2i\sqrt{g}\left(\overline{\theta }\rho ^i\widehat{}_i\theta +i\overline{\theta }\rho _3\theta \right),\rho _3\frac{1}{2}ϵ^{\alpha \beta }\rho _\alpha \rho _\beta =\mathrm{\Gamma }_{04}.$$ Here we introduced the notation $`\rho _\alpha =(\mathrm{\Gamma }_0,\mathrm{\Gamma }_4)`$ ($`\rho _\alpha `$ may be identified with 2-d Dirac matrices times $`I_8`$). Thus the quadratic fermionic part of GS action has exactly the same form as the action for 2-d fermions in curved 2-d space. Assuming the standard $`d^2\sigma \sqrt{g}\overline{\theta }\theta `$ normalization, the corresponding Dirac operator is $$D_F=i\rho ^i\widehat{}_i\rho _3=iw(\mathrm{\Gamma }_0_0+\mathrm{\Gamma }_4_1)\frac{1}{2}i\mathrm{\Gamma }_4\mathrm{\Gamma }_0\mathrm{\Gamma }_4,$$ where the third term came from $`D_0`$. The spectral problem is thus $$[iw(\mathrm{\Gamma }_0_0+\mathrm{\Gamma }_4_1)\frac{1}{2}i\mathrm{\Gamma }_4\mathrm{\Gamma }_0\mathrm{\Gamma }_4]\theta =\lambda \theta .$$ Directly squaring this operator we get $$\left(w^2[(_0\frac{1}{2}w^1\mathrm{\Gamma }_0\mathrm{\Gamma }_4)^2_1^2]+\frac{1}{2}\right)\theta =(\widehat{}^2+\frac{1}{4}R^{(2)}+1)\theta =\lambda ^2\theta .$$ Ignoring the ghosts and longitudinal modes, we are left with a 2-d field theory on $`AdS_2`$ containing five massless scalars, three scalars with mass squared 2, and eight fermions with mass squared 1. Field theories on $`AdS_2`$ were studied in the past (see \[32,,33,,34,,35,,36,,30,,37\]). The fields in the $`𝒩=1`$ scalar supermultiplet in $`AdS_2`$ may have the following bosonic and fermionic masses \[38,,30,,33\]: $$m_B^2=\mu ^2\mu ,m_F=\mu ,$$ where $`\mu `$ is a free parameter. In the case at hand we have 5 “massless” multiplets with $`\mu =1`$ ($`m_B^2=0`$, $`m_F=1`$) and 3 “massive” multiplets with $`\mu =1`$ ($`m_B^2=2`$, $`m_F=1`$). It is possible to combine a $`\mu =1`$ and a $`\mu =1`$ multiplet into an $`𝒩=2`$ multiplet, the dimensional reduction of the 4-d chiral multiplet to 2 dimensions. Two $`\mu =1`$ multiplets also form an $`𝒩=2`$ multiplet which is the dimensional reduction of the 4-d vector multiplet. Three chiral and one vector multiplets in $`D=4`$ make one $`𝒩=4`$ vector in four dimensions, so we conclude that the 8 scalars and 8 fermions that we have found should form one $`𝒩=8`$ multiplet in two dimensions (see, e.g., for related discussion).<sup>7</sup> The mass term in the action (4.1), contains the matrix $`\rho _3=\mathrm{\Gamma }_0\mathrm{\Gamma }_4`$ which has half of its eigenvalues $`1`$ and half $`1`$, i.e. there are actually 4 fermions with $`m_F=1`$ and 4 with $`m_F=1`$, not $`3+5`$. But the sign choice in (4.1) $`m_B^2=\mu ^2\mu `$ rather than $`m_B^2=\mu ^2+\mu `$ is for $`𝒩=1`$ supersymmetry. For extended supersymmetry both signs are possible, so the bosons can be split into 3 and 5 while the fermions are split to 4 and 4. We finally obtain the following partition function with the scalar and spinor Laplace operators defined with respect to the Euclidean $`AdS_2`$ metric with radius 1 ($`R^{(2)}=2`$) $$Z_{B+F}=\frac{\stackrel{8/2}{det}\left(\widehat{}^2+\frac{1}{4}R^{(2)}+1\right)}{det^{3/2}\left(^2+2\right)det^{5/2}\left(^2\right)}.$$ It seems reasonable to impose, as is usually done in discussions of supersymmetric theories in $`AdS_n`$ backgrounds \[40,,41\], proper boundary conditions consistent with supersymmetry. Those imply that the resulting spectra of Laplace operators are discrete in spatial direction (and not continuous as one would normally expect to find in a non-compact hyperbolic space). A direct calculation of the partition function would involve solving the spectral problems (4.1) and $$\left(_0^2_1^2+\frac{m^2}{\sigma ^2}\right)\zeta =\frac{\lambda }{\sigma ^2}\zeta ,$$ with $`m^2=0,2`$. The solutions to the bosonic problem which vanish at $`\sigma =0`$ are $$\zeta (\tau ,\sigma )=e^{ip\tau }\sqrt{\sigma }K_{i\nu }(p\sigma ),\nu ^2=\lambda m^2\frac{1}{4},0\nu <\mathrm{},$$ where $`K_{i\nu }`$ are modified Bessel functions. A calculation of the partition function based on this spectrum is presented in Appendix B. This direct approach suffers from regularization problems, it also does not capture the symmetries of the problem, like supersymmetry. Here we use a different method to evaluate it. 4.2. Vacuum energy Instead of calculating the partition function directly, we could start with the vacuum energy. It is given by the determinant of the operator scaled to remove the factor of $`g^{00}`$ from in front of $`_t^2`$ (see ). Then we use the conformal anomaly, as discussed in Appendix A, to derive the partition function. Let us change the world sheet coordinates so that the $`AdS_2`$ metric is $$ds^2=\frac{1}{\mathrm{cos}^2\rho }\left(dt^2+d\rho ^2\right),\rho [\frac{\pi }{2},\frac{\pi }{2}].$$ The spectra of the Hamiltonians conjugate to this time variable were calculated in :<sup>8</sup> From the group-theoretical point of view, the unitary irreducible representations of the $`AdS_2`$ superalgebra contain: for $`\mu >\frac{1}{2}`$ a scalar field with $`\omega _n^{(B)}=n+\mu `$ and a fermion field with $`\omega _n^{(F)}=n+\mu +\frac{1}{2},`$ and for $`\mu <\frac{1}{2}`$ – a scalar field with $`\omega _n^{(B)}=n+2|\mu |`$ and a fermion field with $`\omega _n^{(F)}=n+|\mu |+\frac{1}{2}.`$ $$\omega _n^{(F)}(\mu )=n+|\mu |+\frac{1}{2},\omega _n^{(F)}(\pm 1)=n+\frac{3}{2},$$ $$\omega _n^{(B)}(\mu )=n+h(\mu ),h(\mu )=\frac{1}{2}(1+\sqrt{1+4m_B^2}),h(1)=2,h(1)=1.$$ Summing over all the modes we get, as in , the 1-loop vacuum energy of this effective 2-d field theory $$E=\frac{1}{2}\underset{n=0}{\overset{\mathrm{}}{}}\left(3\left[\omega _n^{(B)}(1)\omega _n^{(F)}(1)\right]+5\left[\omega _n^{(B)}(1)\omega _n^{(F)}(1)\right]\right).$$ As was extensively discussed in the literature, the properly defined vacuum energy should vanish in the $`AdS`$ case as it does in flat space \[32,,34,,30,,42,,43\] (even though divergences may not cancel out, unless there is a lot of supersymmetry ). However, the direct computation of the sum of the mode energies using $`\zeta `$-function regularization may lead to a non-zero result because the $`\zeta `$-function regularization may not, in general, preserve supersymmetry. Using the standard relations $$\zeta (s,x)\underset{n=0}{\overset{\mathrm{}}{}}(n+x)^s,\zeta (1,x)=\frac{1}{2}\left(x^2x+\frac{1}{6}\right),$$ we find for a boson ($`m_B^2=\mu ^2\mu `$) $$E_B=\frac{1}{2}\underset{n=0}{\overset{\mathrm{}}{}}[n+h(\mu )]=\frac{1}{2}\zeta (1,h(\mu ))=\frac{1}{4}\left(m_B^2+\frac{1}{6}\right),$$ and for a fermion ($`m_F=\mu `$) $$E_F=\frac{1}{2}\underset{n=0}{\overset{\mathrm{}}{}}\left(n+|\mu |+\frac{1}{2}\right)=\frac{1}{2}\zeta (1,|\mu |+\frac{1}{2})=\frac{1}{4}\left(m_F^2\frac{1}{12}\right).$$ In our case we find<sup>9</sup> Note that if we set the mass terms to zero by taking $`\mu =0`$ in (4.1), we get 8 massless scalars and 8 massless fermions in $`AdS_2`$ and their vacuum energies do not cancel – we get $`\frac{1}{2}\times 8[\zeta (1,1)\zeta (1,\frac{1}{2})]=4\times (\frac{1}{8})`$. $$\begin{array}{cc}\hfill E=\frac{1}{4}\left[3\times \left(2+\frac{1}{6}\right)+5\times \frac{1}{6}8\times \left(1\frac{1}{12}\right)\right]=0.& \end{array}$$ The ratio 3:5 of the numbers of the two multiplets is just what is needed for the cancellation. The fact that $`E`$, defined by $`\zeta `$-function regularization, vanishes may be a consequence of the extended $`𝒩=8`$ supersymmetry mentioned above. Indeed, while as was originally suggested for $`AdS_4`$ and confirmed also in \[30,,42,,43\], the $`\zeta `$-function regularization may break supersymmetry and thus may lead to $`E0`$, this does not actually happen in the case of $`N5`$, $`D=4`$ gauged supergravities . The present $`D=2`$ case is thus analogous to those $`D=4`$ cases with large amounts of supersymmetry.<sup>10</sup> The $`E=0`$ property of $`N5`$, $`D=4`$ supergravities might be related to the cancellation of the logarithmic gauge coupling renormalization in these theories (note, in particular, that as was discussed in in the case of the flat space, the vacuum energy as defined by the partition function is the same as the sum of zero-point energies provided $`\zeta (0)=0`$, i.e. if there are no UV infinities). It may seem that the analogy between our $`D=2`$ case and the $`D=4`$ cases is not quite complete since here, in fact, the naive calculation of the coefficient of the logarithmic divergence in terms of the sum of $`\zeta `$-functions gives a nonzero answer (using $`\zeta (0,x)=\frac{1}{2}x`$, we find the total coefficient to be 1). Note, however, that the types of divergences which cancel in the $`D=4`$ cases and do not cancel in the $`D=2`$ case are actually quite different, i.e. the direct comparison is not possible. 4.3. Partition function The vacuum energy calculated in the previous subsection as the sum over oscillator modes corresponds to the determinants of the following (mass $`m`$) bosonic and fermionic spectral problems $$\begin{array}{cc}\hfill \left(_0^2_1^2+\frac{m^2}{\mathrm{cos}^2\rho }\right)\zeta =\lambda _B\zeta ,& \\ \hfill \left(\widehat{}_0^2\widehat{}_1^2+\frac{1}{2}\frac{1}{\mathrm{cos}^2\rho }\right)\theta =\lambda _F^2\theta ,& \end{array}$$ where in the fermionic operator we assume that the covariant derivatives are contracted using flat metric. These are related to the spectral problems (4.1) and (4.1) (apart from the coordinate change) by a rescaling of the right hand side by $`\mathrm{cos}^2\rho `$. As was mentioned above, in curved (e.g., static conformally flat) space the logarithm of the partition function is, in general, different from the vacuum energy defined as a sum over eigen-modes because the time derivative part of the relevant elliptic operators is rescaled by $`g^{00}`$. The determinants of the two operators which differ by such a rescaling are related to each other by a conformal anomaly type equation as discussed in Appendix A. The extra contribution from a mass $`m`$ boson is $$\mathrm{log}det\mathrm{\Delta }_1=\mathrm{log}det\mathrm{\Delta }_M+\frac{1}{4\pi }d^2\sigma \sqrt{g}\left(m^2\mathrm{ln}M+\frac{1}{12}^i\mathrm{ln}M_i\mathrm{ln}M\right),$$ where $`M=\mathrm{cos}^2\rho =1/\sqrt{g}`$. The two terms in the parentheses differ only by a total derivative, but we choose to write it this way to eliminate the boundary terms. Each fermion contributes $$\mathrm{log}det\mathrm{\Delta }_1=\mathrm{log}det\mathrm{\Delta }_{𝒦^2}\frac{1}{4\pi }d^2\sigma \sqrt{g}\left(2\mathrm{ln}𝒦\frac{2}{3}^i\mathrm{ln}𝒦_i\mathrm{ln}𝒦\right),$$ where $`𝒦=\mathrm{cos}\rho =\sqrt{M}`$. Summed together the ‘transverse’ scalars and fermions contribute $$\mathrm{log}Z_{B+F}=\frac{1}{4\pi }d^2\sigma \sqrt{g}\left(\mathrm{ln}M^i\mathrm{ln}M_i\mathrm{ln}M\right).$$ To this we should add the contribution of the ghosts and longitudinal modes. For the ghosts one gets the standard Liouville action $$\mathrm{log}det\mathrm{\Delta }_1^{\mathrm{gh}}=\mathrm{log}det\mathrm{\Delta }_M^{\mathrm{gh}}\frac{26}{12}\times \frac{1}{4\pi }d^2\sigma \sqrt{g}^i\mathrm{ln}M_i\mathrm{ln}M.$$ The longitudinal modes have the same action as the ghosts, but different boundary conditions giving the conformal anomaly $$\mathrm{log}det\mathrm{\Delta }_1^L=\mathrm{log}det\mathrm{\Delta }_M^L+2\times \frac{1}{4\pi }d^2\sigma \sqrt{g}\left(\mathrm{ln}M+\frac{1}{12}^i\mathrm{ln}M_i\mathrm{ln}M\right),$$ so that $$\mathrm{log}Z_{L+\mathrm{gh}}=\frac{1}{4\pi }d^2\sigma \sqrt{g}\left(\mathrm{ln}M^i\mathrm{ln}M_i\mathrm{ln}M\right).$$ Putting it all together we find that the partition function is identically equal to one $$Z_{\mathrm{total}}=\frac{\stackrel{1/2}{det}(\mathrm{\Delta }_{ij}^{\mathrm{gh}}+\delta _{ij})\stackrel{8/2}{det}(\widehat{}^2+\frac{1}{4}R^{(2)}+1)}{det^{1/2}(\mathrm{\Delta }_{ij}+\delta _{ij})det^{3/2}(^2+2)det^{5/2}(^2)}=1.$$ This result is a consequence of the fact that the vacuum energy vanishes, and also of the cancellation of the sum of conformal anomalies for ten bosons with total mass terms 8, eight fermions with 4 times the standard 2-d fermion conformal anomaly and total mass 8, and the conformal gauge ghosts. Note that this is not identical to the conformal anomaly calculation of Section 3.2. Here we did not distinguish between the induced and the fiducial metric. An alternative method of calculating the partition function would be to go back and treat the fiducial metric $`g_{ij}`$ and the induced metric $`h_{ij}`$ as independent. Then only the fiducial metric should be rescaled, while the induced metric should not. It is most convenient to work with flat metric on the strip $$g_{ij}=\delta _{ij},h_{ij}=\frac{1}{\mathrm{cos}^2\rho }\delta _{ij}.$$ That eliminates the problem of the boundary contributions, since the geodesic curvature is zero. This calculation gives the same spectral problem as the vacuum energy calculation for the bosons and ghosts, but not the fermions and longitudinal modes. 5. Circular Wilson loop Another case where the classical solution has an explicit simple form \[47,,25\] is a circular Wilson loop. Like the straight string case, this configuration is useful as it gives a laboratory to investigate many of the issues that arise in the case of the more general bent string configuration. 5.1. Classical solution and quadratic fluctuation action The target space metric in polar coordinates is ($`s=2,3`$)<sup>11</sup> Because of scale invariance, the radius of the circle may be set equal to one. $$ds^2=\frac{1}{w^2}\left(dr^2+r^2d\varphi ^2+dx^sdx^s+dw^2\right)+d\mathrm{\Omega }_5^2.$$ We set $`x^0=\varphi [0,2\pi ]`$ and $`x^1=r[0,1]`$. The classical solution and the induced metric are $$w=\sqrt{1r^2},g_{ij}=\left(\begin{array}{cc}\frac{r^2}{w^2}& 0\\ 0& \frac{1}{w^4}\end{array}\right),\sqrt{g}g^{ij}=\left(\begin{array}{cc}\frac{1}{rw}& 0\\ 0& rw\end{array}\right),$$ $$R^{(2)}=2,$$ i.e. the world sheet metric is again that of $`AdS_2`$ .<sup>12</sup> To put the metric in a more standard form we set $`y=w^1`$. Then $`ds^2=(y^21)^1dy^2+(y^21)d\varphi ^2`$, or in terms of $`\mathrm{tanh}\chi =r`$, $`ds^2=d\chi ^2+\mathrm{sinh}^2\chi d\varphi ^2`$. In the conformal gauge the quadratic part of the bosonic action is (3.1), i.e. $$S=\frac{1}{2}d^2\sigma \sqrt{g}\left(g^{ij}D_i\zeta ^aD_j\zeta ^a+X_{ab}\zeta ^a\zeta ^b+g^{ij}D_i\zeta ^qD_j\zeta ^q\right),$$ where in the present case $$X^{ab}=2\delta ^{ab}g^{ij}\eta _i^a\eta _j^b,\eta _0=(\frac{r}{w},0,0,0,0),\eta _1=(0,\frac{1}{w},0,0,\frac{r}{w^2}).$$ The nonzero components of the spin connection in the target space are $$\mathrm{\Omega }_0^{01}=1,\mathrm{\Omega }_0^{04}=\frac{r}{w},\mathrm{\Omega }_1^{14}=\mathrm{\Omega }_2^{24}=\mathrm{\Omega }_3^{34}=\frac{1}{w},$$ so that all of the covariant derivatives are trivial, $`D_i=_i`$, except for $$D_0\zeta ^0=_0\zeta ^0+\zeta ^1\frac{r}{w}\zeta ^4,D_0\zeta ^1=_0\zeta ^1\zeta ^0,D_0\zeta ^4=_0\zeta ^4+\frac{r}{w}\zeta ^0,$$ $$\begin{array}{cc}\hfill D_1\zeta ^1=_1\zeta ^1\frac{1}{w}\zeta ^4,D_1\zeta ^4=_1\zeta ^4+\frac{1}{w}\zeta ^1.& \end{array}$$ The covariant derivative $`_i`$ in the ghost action $$S=\frac{1}{2}d^2\sigma \sqrt{g}\left(g^{ij}_iϵ^\alpha _jϵ^\alpha \frac{1}{2}R^{(2)}ϵ^\alpha ϵ^\alpha \right),$$ includes the world-sheet spin connection, whose only nonzero component is $`\omega _0^{01}=w^1`$. This derivative is not the same as the covariant derivative in (5.1). However, if we rotate the fields $$\left(\begin{array}{c}\stackrel{~}{\zeta }^1\\ \stackrel{~}{\zeta }^4\end{array}\right)=\left(\begin{array}{cc}\mathrm{cos}\alpha & \mathrm{sin}\alpha \\ \mathrm{sin}\alpha & \mathrm{cos}\alpha \end{array}\right)\left(\begin{array}{c}\zeta ^1\\ \zeta ^4\end{array}\right),\mathrm{cos}\alpha =w,\mathrm{sin}\alpha =r,\frac{d\alpha }{dr}=\frac{1}{w},$$ the mass matrix becomes diagonal, $`\stackrel{~}{X}_{ab}=\mathrm{diag}(1,1,2,2,2)`$, and the only nontrivial covariant derivatives are $`D_0\zeta ^0=_0\zeta ^0+w^1\stackrel{~}{\zeta }^1`$ and $`D_0\stackrel{~}{\zeta }^1=_0\stackrel{~}{\zeta }^1w^1\zeta ^0`$. Then the longitudinal modes $`\zeta ^0`$ and $`\stackrel{~}{\zeta }^1`$ again have the same action as the ghosts, leaving us with three massive and five massless transverse oscillations. The same conclusion is reached by starting with the Nambu form of the action and choosing static gauge, where $`r`$ and $`\varphi `$ in (5.1) are identified with the world-sheet coordinates. Let us denote $`\xi ^s,\xi ^4,\xi ^q`$ the fluctuations of the $`x^s`$, $`w`$ and the $`S^5`$ coordinates respectively. After rescaling $`\overline{\zeta }^s=w^1\xi ^s`$, the action is (here $`g_{ij}`$ is the induced metric) $$S=\frac{1}{2}d^2\sigma \sqrt{g}\left(g^{ij}_i\overline{\zeta }^s_j\overline{\zeta }^s+2\overline{\zeta }^s\overline{\zeta }^s+g^{ij}_i\xi ^4_j\xi ^4+2\xi ^4\xi ^4+g^{ij}_i\xi ^q_j\xi ^q\right),$$ and the fields are normalized as $$\xi ^2=𝑑r𝑑\varphi \sqrt{g}\left(\overline{\zeta }^s\overline{\zeta }^s+\frac{1}{w^2}\xi ^4\xi ^4+\xi ^q\xi ^q\right).$$ Note that the field $`\xi ^4`$ (and $`\zeta ^4`$ above) is not normal to the surface, but $`\stackrel{~}{\zeta }^4`$ is. As explained in Section 3.3, this choice of gauge has a non trivial ghost determinant (3.1). In the present case $`h_{ij}^{}=\mathrm{diag}(1/w^2,r^2/w^2)`$, so that $$\mathrm{\Delta }_{\mathrm{gh}}=\stackrel{1/2}{det}(h_{ik}^{}h^{kj})=\stackrel{1/2}{det}w^2.$$ The most natural way to regularize this determinant is by redefining the norm of the $`\xi ^4`$, thus removing the extra normalization factor in (5.1). Then the result for the partition function in this gauge will be identical to the conformal gauge expression apart from the contributions of the ghosts and the longitudinal modes.<sup>13</sup> If we were to expand the action without fixing the gauge $`\zeta ^0=\zeta ^1=0`$, we would find the same action, but with $`\xi ^4`$ replaced by $`\stackrel{~}{\zeta }^4`$, with the canonical normalization. The two longitudinal fluctuations $`\zeta ^0`$ and $`\stackrel{~}{\zeta }^1`$ drop out of the action. Then one could choose the gauge $`\zeta ^0=\stackrel{~}{\zeta }^1=0`$ (this is the gauge used in \[7,,10\] in the context of the bent string configuration). $`\stackrel{~}{\zeta }^1`$ and $`\zeta ^1`$ are related through a rotation by an angle $`\mathrm{cos}\alpha =w`$. This rotation introduces a Jacobian which exactly cancels the ghost determinant in the former gauge. It is also easy to show directly that the ghost determinant is trivial in this gauge. Before gauge fixing, the fermionic Lagrangian is (3.1) $$L_{2\mathrm{F}}=i\left(\sqrt{g}g^{ij}\delta ^{IJ}ϵ^{ij}s^{IJ}\right)\overline{\theta }^I\rho _iD_j\theta ^J.$$ To put this action in a 2-d covariant fermionic form in terms of the zweibein and spin connection of the induced metric we apply a local $`SO(1,9)`$ rotation to transform the projected Dirac matrices $`\rho _i`$ into constant Dirac matrices contracted with the induced zweibein as discussed in section 3.4. We get (cf. (3.1),(3.1)) $$\begin{array}{cc}\hfill D_i\theta ^J& =𝒟_i\theta ^J\frac{i}{2}ϵ^{JK}\rho _i\theta ^K,\hfill \\ \hfill \rho _0& =\eta _0^a\mathrm{\Gamma }_a=\frac{r}{w}\mathrm{\Gamma }_0=e_0^\alpha S\mathrm{\Gamma }_\alpha S^1,\hfill \\ \hfill \rho _1& =\eta _1^a\mathrm{\Gamma }_a=\frac{1}{w}\mathrm{\Gamma }_1\frac{r}{w^2}\mathrm{\Gamma }_4=e_1^\alpha S\mathrm{\Gamma }_\alpha S^1,\hfill \\ \hfill 𝒟_0& =\left(_0+\frac{1}{2}\mathrm{\Gamma }_{01}\frac{r}{2w}\mathrm{\Gamma }_{04}\right)=S\widehat{}_0S^1,\hfill \\ \hfill 𝒟_1& =\left(_1\frac{1}{2w}\mathrm{\Gamma }_{14}\right)=S\widehat{}_1S^1.\hfill \end{array}$$ The rotation matrix is $$S=\mathrm{exp}\left(\frac{\alpha }{2}\mathrm{\Gamma }_{14}\right),$$ with the same angle $`\alpha `$ as in (5.1). $`\widehat{}_i`$ is the covariant derivative with spinor world-sheet connection, $$\widehat{}_0=_0+\frac{1}{2w}\mathrm{\Gamma }_{01},\widehat{}_1=_1.$$ It is therefore natural to transform $`\theta ^I`$ to the new variable $`\mathrm{\Psi }^I`$ $$\theta ^I=S\mathrm{\Psi }^I.$$ Choosing the gauge $`\mathrm{\Psi }^1=\mathrm{\Psi }^2`$, the fermionic Lagrangian becomes $$\begin{array}{cc}\hfill L_{2\mathrm{F}}& =2i\sqrt{g}\left(g^{ij}\overline{\mathrm{\Psi }}e_i^\alpha \mathrm{\Gamma }_\alpha _j\mathrm{\Psi }+i\overline{\mathrm{\Psi }}\mathrm{\Gamma }_{01}\mathrm{\Psi }\right)\hfill \\ & =2i\overline{\mathrm{\Psi }}\left(\frac{1}{w^2}\mathrm{\Gamma }_0_0+\frac{r}{w}\mathrm{\Gamma }_1_1+i\frac{r}{w^3}\mathrm{\Gamma }_{01}\right)\mathrm{\Psi }.\hfill \end{array}$$ Here the 10-d Dirac matrices $`\mathrm{\Gamma }_\alpha `$ play the role of world sheet Dirac matrices, as we can choose a representation in terms of the Pauli matrices $`\mathrm{\Gamma }_0=i\sigma _2\times I_8`$$`\mathrm{\Gamma }_1=\sigma _1\times I_8`$. As in the case of the straight string (4.1), this is the action for a spinor of mass $`\pm 1`$ in $`AdS_2`$. Another natural way to fix the $`\kappa `$ symmetry used in \[48,,6\]<sup>14</sup> $`\mathrm{\Gamma }_{0123}=i\gamma _4\times I_4\times I_2`$ in the notation of , where 10-d Dirac matrices are represented as $`\mathrm{\Gamma }^a=\gamma ^a\times I_4\times \sigma _1`$, $`a=0,1,2,3,4`$. $$\theta ^1=\mathrm{\Gamma }_{0123}\theta ^2,\mathrm{i}.\mathrm{e}.\theta ^1=i\mathrm{\Gamma }_4\theta ^2.$$ This gauge leads to the same result for the action as the $`\theta ^1=\theta ^2`$ gauge as we shall explain below.<sup>15</sup> Note that in the straight string case (4.1) this gauge is degenerate. Expressing $`\theta ^2`$ in terms of $`\theta ^1\theta `$ one can check that $$D_i\theta ^1=\frac{1}{\sqrt{w}}\left(_i+\frac{1}{2}\mathrm{\Gamma }_{i1}\right)\left(\sqrt{w}\theta \right),$$ and , in terms of $$\vartheta \sqrt{w}\theta ,$$ the Lagrangian is $$L_{2\mathrm{F}}=2i\overline{\vartheta }\left[\left(\frac{1}{w^3}\mathrm{\Gamma }_0\frac{r}{w^3}\mathrm{\Gamma }_4\right)\left(_0+\frac{1}{2}\mathrm{\Gamma }_{01}\right)+\frac{r}{w}\mathrm{\Gamma }_1_1\right]\vartheta .$$ To simplify this expression, we again use a rotation, this time in the 0-4 plane. Define $$\psi =\mathrm{exp}\left(\frac{\beta }{2}\mathrm{\Gamma }_{04}\right)\vartheta ,$$ where $$\mathrm{cosh}\beta =\frac{1}{w},\mathrm{sinh}\beta =\frac{r}{w},\frac{d\beta }{dr}=\frac{1}{w^2}.$$ Then $$L_{2\mathrm{F}}=2i\overline{\psi }\left[\frac{1}{w^2}\mathrm{\Gamma }_0\left(_0+\frac{1}{2}\mathrm{\Gamma }_{01}\right)+\frac{r}{w}\mathrm{\Gamma }_1_1\frac{r}{w^3}\mathrm{\Gamma }_{104}\right]\psi .$$ Though this action looks different from (5.1), it also describes a fermion of mass $`\pm `$1 (the mass term $`\mathrm{\Gamma }_{104}=i\mathrm{\Gamma }_{23}`$ commutes with $`\mathrm{\Gamma }_0`$ and $`\mathrm{\Gamma }_1`$, but is antihermitian, and its square is 1). The normalization of $`\psi `$ is $$\psi ^2=𝑑r𝑑\varphi \sqrt{g}w^1\overline{\psi }\psi ,$$ which is different from normalization of $`\mathrm{\Psi }`$ in (5.1). Like in the bosonic case, the difference of the normalizations of the fields in the two $`\kappa `$-symmetry gauges may be attributed to the difference in the corresponding ghost determinants. Indeed, if the gauge condition is $`\theta ^1=H\theta ^2`$ where $`H`$ is some matrix ($`H=1`$ and $`H=i\mathrm{\Gamma }_4`$ in the two gauges discussed above), then it is easy to find the ghost determinant corresponding to the transformation (3.1). In the cases we are interested in the case where the $`\overline{x}`$ background is constant on $`S^5`$ (i.e. when $`\stackrel{~}{\rho }_i=\rho _i`$) $$\delta _\kappa \theta ^1=\rho _i^{}k^i,\delta _\kappa \theta ^2=\rho _i^+k^i,$$ where $`\rho _i^\pm =(g_{ij}\pm e_{ij})\rho _j`$, where $`e^{ij}=\frac{ϵ^{ij}}{\sqrt{g}}`$ and $`k_i`$ are unconstrained vector-spinor parameters, normalized as $`k_i^2=d^2\sigma \sqrt{g}g^{ij}\overline{k}_ik_j`$. Then the ghost determinant is the inverse square root of the determinant of the spinor matrix<sup>16</sup> The ghost determinant can be obtained from the path integral $$\mathrm{\Delta }_{\mathrm{gh}}[dk_i]\mathrm{exp}\left[d^2\sigma \sqrt{g}\overline{k}^i(\rho _i^{}\rho _i^+H^{})(\rho _j^{}H\rho _j^+)k_j\right]=1.$$ $$g^{ij}(\rho _i^{}\rho _i^+H^{})(\rho _j^{}H\rho _j^+).$$ Since $`\rho _{(i}\rho _{j)}=g_{ij}`$ this matrix is a trivial constant in the $`\theta ^1=\theta ^2`$ gauge when $`H=1`$, but, in general, it will depend on the components of the metric (i.e. on the coordinates) when $`H1`$. In the gauge $`\theta ^1=i\mathrm{\Gamma }_4\theta ^2`$ the resulting local ghost determinant should “compensate” (in an appropriate regularization scheme) for the difference in normalizations of the spinors $`\mathrm{\Psi }`$ in (5.1) and $`\psi `$ in (5.1), explaining the equivalence of the results in the two $`\kappa `$-symmetry gauges. 5.2. Partition function We thus end up with the same partition function (4.1) as in the straight string case, i.e. with that of a supersymmetric field theory on $`AdS_2`$. The result is again a constant whose precise value depends on regularization and measures for the fields. This should not come as a surprise, since the circle and the straight line are related by a special conformal transformation, and the minimal surfaces also transform into each other. This is not to say that the partition functions should be identical, there is a subtle difference. Indeed, already the classical actions for a circle and a straight line are different. The reason can be traced to the inclusion of the point at infinity. The same subtlety should be present at the level of 1-loop partition function. In the case of the straight line it is natural to work with the strip model for $`AdS_2`$, while for the circle, the Poincaré disk is more natural. In calculating the determinant for the former we should include also functions that do not behave well at infinity, while in the circle case those should not be included. It is therefore probable that the calculation in Appendix B is more appropriate for this problem rather than the straight string case. 6. ‘Parallel Lines’ Our main interest is in the minimal surface ending at the boundary of $`AdS_5\times S^5`$ which is related to the correlation function of two anti-parallel Wilson loops. The minimal surface was constructed in and accounts for the leading large $`\lambda `$ behavior ($`\frac{c_0\sqrt{\lambda }}{L}`$) of the “quark – anti-quark” (W-boson) potential in $`𝒩=4`$ SYM theory. The first correction $`\frac{c_1}{L}`$ to the potential will be given by the one-loop partition function of the type we study here. While some aspects of this computation were addressed before \[6,,7,,10\], our aim below will be to clarify some previously encountered problems and to set up a systematic framework which should allow to compute the finite numerical coefficient $`c_1`$. 6.1. The classical solution In this section we will write the $`AdS_5\times S^5`$ metric in terms of the coordinate $`y=w^1`$ (cf. (3.1)) $$ds^2=R^2(y^2dx^ndx^n+\frac{dy^2}{y^2}+d\mathrm{\Omega }_5^2).$$ Here $`n=0,1,2,3`$ and we will use the index 4 to label the coordinate $`y`$. We will often set the radius $`R`$ to be 1 in what follows.<sup>17</sup> To make the flat space limit explicit one should define the coordinate $`\phi `$ related to $`y`$ by $`y=R^1e^{\phi /R}`$. Then $`ds^2=e^{2\phi /R}dx^ndx^n+d\phi ^2+R^2d\mathrm{\Omega }_5`$ which becomes flat in the $`R\mathrm{}`$ limit. If the Wilson lines are extended in the $`x^0`$ direction and located at $`x^1=\pm \frac{L}{2}`$, the minimal surface is given by a function $`y(x^1)`$ (we use world-sheet coordinates $`\sigma ^i=x^i=(\tau ,\sigma )`$, $`0<\tau <T`$). The bosonic part of string action is then ($`y^{}_1y`$) $$S=\frac{R^2}{2\pi \alpha ^{}}T𝑑\sigma \sqrt{y^2+y^4}.$$ The stationary point is determined by the second-order equation $`yy^{\prime \prime }=4y^2+2y^4`$ with the first integral being $$y^2=\frac{y^8}{y_0^4}y^4.$$ $`y_0`$ is an integration constant, the minimal value of $`y`$. The special case of $`y_0=0`$ corresponds to the “straight string” configuration discussed in Section 4. This special solution is a useful reference point: near the boundaries of the $`\sigma `$-interval it gives a good approximation to the general solution. The induced metric is $$g_{ij}=\left(\begin{array}{cc}y^2& 0\\ 0& y^6\end{array}\right),\sqrt{g}=y^4,\sqrt{g}g^{ij}=\left(\begin{array}{cc}y^2& 0\\ 0& y^2\end{array}\right).$$ The distance between the quark and anti-quark $`L`$ is related to $`y_0`$ by $$L=_{L/2}^{L/2}𝑑\sigma =2_{y_0}^{\mathrm{}}\frac{dy}{y^2\sqrt{\frac{y^4}{y_0^4}1}}=\frac{\kappa _0}{y_0},$$ $$\kappa _0\frac{(2\pi )^{3/2}}{[\mathrm{\Gamma }(\frac{1}{4})]^2}.$$ We shall often set $`y_0=1`$ as the dependence on this parameter can be easily restored by rescalings ($`\tau y_0^1\tau `$, i.e. $`T\frac{T}{y_0^2}`$). Then $$y^2=y^8y^4,y^{\prime \prime }=4y^72y^3.$$ Let us first review how the classical contribution is computed. The action (6.1) takes the following value on the solution $$S=\frac{R^2T}{2\pi \alpha ^{}y_0^2}_{L/2}^{L/2}𝑑\sigma y^4.$$ Since $`(y^3y^{})^{}=\frac{y^4+y_0^4}{y_0^2}`$ is a total derivative (which goes to the boundary where $`y=\mathrm{}`$ and gives only a trivial divergence), we can replace $`y^4`$ by $`y_0^4`$, assuming that the infinite boundary contribution should be dropped. This prescription is the same as normalizing the partition function to the straight line case (and is essentially equivalent to the one in : the Legendre transform subtracts the same boundary term or total derivative<sup>18</sup> This is an example of an amusing relation. The Legendre transform can be written as an integral over a (rather complicated) total derivative. Instead, one can note that for smooth loops the Legendre transform which is equal to the divergence in the area is also equal (asymptotically) to the geodesic curvature $`K`$. Then we can use the Gauss-Bonnet theorem to write the action as $`S=\frac{R^2}{2\pi \alpha ^{}}d^2\sigma \sqrt{g}(1+\frac{1}{2}R^{(2)})\frac{R^2}{\alpha ^{}}\chi `$, where $`\chi `$ is the Euler number. Since $`R^{(2)}`$ approaches $`2`$, the integral is manifestly convergent. For the present geometry $`\chi =0`$. ). Then $$S=\frac{R^2T}{2\pi \alpha ^{}}y_0^2_{L/2}^{L/2}𝑑\sigma =\frac{R^2TL}{2\pi \alpha ^{}}y_0^2=\frac{\lambda ^{1/2}(2\pi )^2}{[\mathrm{\Gamma }(\frac{1}{4})]^4}\frac{T}{L},$$ where $`\lambda =4\pi g_sN=\frac{R^4}{\alpha ^2}`$. This result is the same as in , here found in ‘one shot’ (without doing any further integrals). In the flat space limit ($`R\mathrm{}`$) one finds that the quantum correction (2.1) to the potential vanishes because of the cancellation of the bosonic and fermionic contributions due to effective 2-d supersymmetry present after gauge fixing (see Section 2.3). As was pointed out in , the 1-loop $`\frac{c_1}{L}`$ correction to the effective potential may not, however, vanish in the present curved space case as there is no reason to expect that the action expanded near the solution $`y=y(\sigma )\mathrm{const}`$ should have an effective world-sheet supersymmetry. 6.2. Quadratic fluctuations: bosons In conformal gauge the bosonic action is (3.1), where $`X_{pq}=0`$ and<sup>19</sup> In this form of the metric (6.1) the vielbein components are $`E_n^m=y\delta _n^m`$, $`E_4^4=y^1`$. If we restore the $`R`$ dependence, $`X_{ab}\frac{1}{R^2}X_{ab}`$, then the “mass term” vanishes in the flat space limit as it should. $$\begin{array}{cc}& X^{ab}=2\delta ^{ab}g^{ij}\eta _i^a\eta _j^b,\eta _0^a=(y,0,0,0,0),\eta _1^a=(0,y,0,0,y^1y^{}),\hfill \\ & D_i\zeta ^a=_i\zeta ^a+w_i^{ab}\zeta ^b,w_i^{ab}=_ix^\mu \mathrm{\Omega }_\mu ^{ab},w_i^{a4}=y\delta _i^a=w_i^{4a}.\hfill \end{array}$$ The ghost action is (3.1) or (5.1) where the covariant derivative $`_i`$ includes the world-sheet spin connection $`\omega _0^{01}=y^3y^{}`$. The action contains the curvature $`R^{(2)}`$ of the induced metric $`g_{ij}`$ $$\sqrt{g}R^{(2)}=\left(\frac{1}{y^2}\right)^{\prime \prime },R^{(2)}=2\left(1+\frac{1}{y^4}\right).$$ Unlike the circle case, here there is no obvious rotation of the fields $`\zeta ^a`$ such that the contribution of the longitudinal modes becomes the same as that of the ghosts.<sup>20</sup> In fact, the eigenvalues of the mass matrix $`X^{ab}`$ are $`(1,1,2,2,2)`$, not $`(\frac{1}{2}R^{(2)},\frac{1}{2}R^{(2)},4+R^{(2)},2,2)`$. In addition there are extra connection terms that remain after the rotation. Choosing the static gauge in the Nambu action we denote by $`\xi ^s`$ ($`s=2,3`$) the fluctuations of the two “longitudinal” 3-brane directions, by $`\xi ^4`$ the fluctuation along the radial $`y`$-direction, and by $`\xi ^q`$ ($`q=5,\mathrm{},9`$) the fluctuations in the 5-sphere directions. Their natural norms are $$\xi ^2=d^2\sigma \sqrt{g}(y^2\xi ^s\xi ^s+y^2\xi ^4\xi ^4+\xi ^q\xi ^q).$$ Introducing the rescaled fields<sup>21</sup> It should be noted that redefinitions we make are accompanied by Jacobians and thus do not introduce new quadratic or linear divergences. $$\zeta ^s=y\xi ^s,\stackrel{~}{\zeta }^4=y^3\xi ^4,$$ one finds (after integration by parts and use of the properties of the classical background (6.1)) the following expression for the quadratic fluctuation part of the gauge-fixed action<sup>22</sup> As before, we absorb $`\frac{1}{2\pi \alpha ^{}}`$ factor in the action into a redefinition of the fluctuation fields. $$S_{2\mathrm{B}}=\frac{1}{2}d^2\sigma \sqrt{g}\left[g^{ij}_i\zeta ^s_j\zeta ^s+2\zeta ^s\zeta ^s+g^{ij}_i\stackrel{~}{\zeta }^4_j\stackrel{~}{\zeta }^4+(R^{(2)}+4)\stackrel{~}{\zeta }^4\stackrel{~}{\zeta }^4+g^{ij}_i\xi ^q_j\xi ^q\right].$$ As follows from (6.1), the fields in (6.1) are normalized as follows $$\xi ^2=d^2\sigma \sqrt{g}(\zeta ^s\zeta ^s+y^4\stackrel{~}{\zeta }^4\stackrel{~}{\zeta }^4+\xi ^q\xi ^q).$$ Thus, while the action $`S_{2\mathrm{B}}`$ seems to have a ‘covariant’ 2-d form with respect to the induced geometry (two massive scalars, one scalar with a potential, and 5 massless scalars in external 2-d metric), this is not true for the measure because of the $`y^4`$ factor in the $`\stackrel{~}{\zeta }^4\stackrel{~}{\zeta }^4`$ part. This is remedied by the inclusion of the ghost determinant (3.1), where in our case $`h_{ij}^{}=\mathrm{diag}(y^2,y^2)`$, so that $$\mathrm{\Delta }_{\mathrm{gh}}=\stackrel{1/2}{det}(1/y^4).$$ The most natural regularization of this determinant is achieved by rescaling the field $`\stackrel{~}{\zeta }^4`$, which cancels the renormalization factor in (6.1), much as in the case of the circle in Section 5.1 (see (5.1)). The same expression for the action (6.1) was found in . There instead of $`\stackrel{~}{\zeta }^4`$ the authors used the fluctuating field normal to the surface $`\eta ^4\mathrm{sin}\alpha \zeta ^1+\mathrm{cos}\alpha \zeta ^4=y^3y^{}\xi ^1+y^3\xi ^4,`$ where $`\alpha `$ is the angle defined by $`\mathrm{cos}\alpha =y^2,\alpha ^{}=2y`$, and $`\zeta ^a=E_\mu ^a\xi ^\mu `$ are the target-space vielbein components of $`\xi ^\mu `$. The normalization of $`\eta ^4`$ is canonical, and it is easy to see that the ghost determinant is trivial in that gauge. 6.3. Quadratic fluctuations: fermions In the present case of the classical solution (6.1), which is constant in $`S^5`$ directions, one finds that the leading quadratic part of the fermionic action given by (3.1) depends on $$\rho _0=y\mathrm{\Gamma }_0,\rho _1=y\mathrm{\Gamma }_1+y^1y^{}\mathrm{\Gamma }_4,𝒟_i=_i+\frac{1}{2}y\mathrm{\Gamma }_i\mathrm{\Gamma }_4,$$ where we used the fact that for $`AdS_5`$ space the non-vanishing components of the connection are $`\mathrm{\Omega }_\mu ^{n4}=E_\mu ^n,n=0,1,2,3`$. Then (3.1) becomes $$L_{2\mathrm{F}}=i\left(\sqrt{g}g^{ij}\delta ^{IJ}ϵ^{ij}s^{IJ}\right)\left(\overline{\theta }^I\rho _i𝒟_j\theta ^J\frac{1}{2}iϵ^{JK}\overline{\theta }^I\rho _i\rho _j\theta ^K\right).$$ Here $`g_{ij}`$ is the Minkowski version of the induced metric (6.1), i.e. the corresponding 2-d vielbein components $`e_i^\alpha `$ are $$e_0^0=y,e_1^1=y^3,g_{ij}=\mathrm{diag}(y^2,y^6).$$ The crucial observation, that allows us to put the action (6.1) into a simple 2-d covariant form, is that the combination of $`\mathrm{\Gamma }_1`$ and $`\mathrm{\Gamma }_4`$ which appears in $`\rho _1`$ can be interpreted as a (local, $`\sigma `$-dependent) rotation of $`\mathrm{\Gamma }_1`$ in the 1-4 plane $$S\mathrm{\Gamma }_1S^1=\mathrm{cos}\alpha \mathrm{\Gamma }_1+\mathrm{sin}\alpha \mathrm{\Gamma }_4=y^2\mathrm{\Gamma }_1+y^4y^{}\mathrm{\Gamma }_4=y^3\rho _1,$$ where $$S=\mathrm{exp}\left(\frac{\alpha }{2}\mathrm{\Gamma }_1\mathrm{\Gamma }_4\right),\mathrm{cos}\alpha =y^2,\mathrm{sin}\alpha =y^4y^{},\alpha ^{}\frac{d\alpha }{d\sigma }=2y.$$ Making the field redefinition $$\theta ^I\mathrm{\Psi }^IS^1\theta ^I,$$ we then find that (6.1) takes the following simple ‘2-d covariant’ form $$L_{2\mathrm{F}}=i\left(\sqrt{g}g^{ij}\delta ^{IJ}ϵ^{ij}s^{IJ}\right)\left(\overline{\mathrm{\Psi }}^I\tau _i\widehat{}_j\mathrm{\Psi }^J\frac{1}{2}iϵ^{JK}\overline{\mathrm{\Psi }}^I\tau _i\tau _j\mathrm{\Psi }^K\right),$$ where $`\tau _i`$ play the role of the curved space 2-d Dirac matrices<sup>23</sup> To prove eq.(6.1) one should note that $`\tau _i=S^1\rho _iS`$ and that $`\stackrel{~}{𝒟}_iS^1𝒟_iS`$ is found to be (see (6.1)) $`\stackrel{~}{𝒟}_0=_0+\frac{1}{2}y\mathrm{\Gamma }_0(y^4y^{}\mathrm{\Gamma }_1+y^2\mathrm{\Gamma }_4)=\widehat{}_0+B_0`$, $`\stackrel{~}{𝒟}_1=_1y\mathrm{\Gamma }_1\mathrm{\Gamma }_4+\frac{1}{2}y\mathrm{\Gamma }_1\mathrm{\Gamma }_4=_1+B_1`$, where $`B_0=\frac{1}{2}y^2\tau _0\mathrm{\Gamma }_4`$, $`B_1=\frac{1}{2}y^2\tau _1\mathrm{\Gamma }_4`$. Finally, one observes that the connection $`B_i`$ drops out from the action since $`\tau ^iB_i=0`$$`ϵ^{ij}\tau _iB_j=0`$. $$\tau _i=e_i^\alpha \mathrm{\Gamma }_\alpha ,\tau _0=S^1\rho _0S=y\mathrm{\Gamma }_0,\tau _1=S^1\rho _1S=y^3\mathrm{\Gamma }_1,$$ and $`\widehat{}_i`$ is the 2-d curved space spinor covariant derivative corresponding to (6.1) $$\widehat{}_k=_k+\frac{1}{4}\omega _k^{\alpha \beta }\mathrm{\Gamma }_{\alpha \beta },\widehat{}_0=_0+\frac{1}{2}y^3y^{}\mathrm{\Gamma }_0\mathrm{\Gamma }_1,\widehat{}_1=_1.$$ The Lagrangian (6.1) is then $$L_{2\mathrm{F}}=i\left(\sqrt{g}g^{ij}ϵ^{ij}\right)\overline{\mathrm{\Psi }}^1\tau _i\widehat{}_j\mathrm{\Psi }^1+i\left(\sqrt{g}g^{ij}+ϵ^{ij}\right)\overline{\mathrm{\Psi }}^2\tau _i\widehat{}_j\mathrm{\Psi }^2ϵ^{ij}\overline{\mathrm{\Psi }}^1\tau _i\tau _j\mathrm{\Psi }^2.$$ It is easy to see that the covariant derivative and ‘mass’ terms here are separately invariant under the leading-order $`\kappa `$-symmetry transformations (see (3.1), (5.1)) $`\delta _\kappa \theta ^I=\rho _i\kappa ^{iI}`$ or their ‘rotated’ form $$\delta _\kappa \mathrm{\Psi }^I=\tau _i\kappa ^{iI},\kappa ^{iI}=S^1k^{iI}.$$ Fixing the $`\kappa `$-symmetry gauge by the condition $$\theta ^1=\theta ^2,\mathrm{i}.\mathrm{e}.\mathrm{\Psi }^1=\mathrm{\Psi }^2\mathrm{\Psi },$$ we get $$L_{2\mathrm{F}}=2i\sqrt{g}\left(\overline{\mathrm{\Psi }}\tau ^i\widehat{}_i\mathrm{\Psi }+i\overline{\mathrm{\Psi }}\tau _3\mathrm{\Psi }\right),$$ $$\tau _3\frac{ϵ^{ij}}{2\sqrt{g}}\tau _i\tau _j=\mathrm{\Gamma }_0\mathrm{\Gamma }_1,(\tau _3)^2=1.$$ Note that choosing this gauge before the rotation, the action (6.1) may be written as $$L_{2\mathrm{F}}=2i\sqrt{g}\overline{\theta }\rho ^i𝐃_i\theta =i\sqrt{g}(\overline{\theta }\rho ^i_i\theta _i\overline{\theta }\rho ^i\theta +ie_{ij}\overline{\theta }\rho ^i\rho ^j\theta ),$$ where $`𝐃_j=𝒟_j+\frac{1}{2}ie_{jk}\rho ^k,`$ $`\rho ^i=g^{ij}\rho _j`$, $`e_{jk}\frac{1}{\sqrt{g}}g_{jj^{}}g_{kk^{}}ϵ^{j^{}k^{}}`$. To see that the rotation (6.1),(6.1) is indeed a special case of the general rotation (3.1) discussed in Section 3.4 we note that here the tangent $`t_\alpha ^\mu `$ and normal $`n^\mu `$ vector components are (here $`\mu ,\nu =0,1,4`$ label the target space $`AdS_3`$ coordinates inside $`AdS_5`$, i.e. $`G_{\mu \nu }=(y^2,y^2,y^2`$)): $$t_{\widehat{0}}^\mu =(1,0,0),t_{\widehat{1}}^\mu =(0,y^3,y^3y^{}),n^\mu =(0,y^5y^{},y^1).$$ Then $`\rho _\alpha =(\rho _{\widehat{0}},\rho _{\widehat{1}})`$ and $`\rho _s\rho _{}`$ are (cf. (6.1)) $$\rho _{\widehat{0}}=\mathrm{\Gamma }_0,\rho _{\widehat{1}}=y^3\rho _1=y^2\mathrm{\Gamma }_1+y^4y^{}\mathrm{\Gamma }_4,\rho _{}=y^3\rho _1=y^4y^{}\mathrm{\Gamma }_4+y^2\mathrm{\Gamma }_4,$$ so that $$S^1dS=\frac{1}{4}(\rho _\alpha d\rho ^\alpha +\rho _{}d\rho _{})=\frac{1}{2}\mathrm{\Gamma }_1\mathrm{\Gamma }_4d\alpha =y\mathrm{\Gamma }_1\mathrm{\Gamma }_4d\sigma ,$$ in agreement with (6.1). This is $`U(1)`$ rotation that does not lead to a non-trivial Jacobian in the present case of non-chiral 2-d fermions. Another possible gauge choice is the analog of the covariantized light cone gauge of (see (3.1),(6.1)): $$(\rho _0+\rho _1)\mathrm{\Psi }^1=0,(\rho _0\rho _1)\mathrm{\Psi }^2=0,$$ or, explicitly, after the rotation (6.1),<sup>24</sup> Note that this gauge choice is different from the one in where instead of $`\tau _\pm `$ the combinations $`\mathrm{\Gamma }_0\pm \mathrm{\Gamma }_1^{}`$ ($`\mathrm{\Gamma }_1^{}=S\mathrm{\Gamma }_1S^1`$) were used (the rotation was not explicitly done in ). $$\tau _+\mathrm{\Psi }^1=(y\mathrm{\Gamma }_0+y^3\mathrm{\Gamma }_1)\mathrm{\Psi }^1=0,\tau _{}\mathrm{\Psi }^2=(y\mathrm{\Gamma }_0y^3\mathrm{\Gamma }_1)\mathrm{\Psi }^2=0.$$ The resulting action is essentially the same as (6.1) with left and right parts of $`\mathrm{\Psi }`$ explicitly separated. Choosing a representation for $`\mathrm{\Gamma }_a`$ such that $`\mathrm{\Gamma }_{0,1}`$ are 2-d Dirac matrices times a unit $`8\times 8`$ matrix, i.e.<sup>25</sup> Recall that the original $`32\times 32`$ Dirac matrices are such that $`\mathrm{\Gamma }_a=\gamma _a\times I_4\times \sigma _1`$, or simply $`\mathrm{\Gamma }_a=\gamma _a\times I_4`$ on a 16-subspace of left MW spinors, with $`\gamma _{(a}\gamma _{b)}=\mathrm{diag}(1,+1,+1,+1,+1)`$. We are not distinguishing between $`\mathrm{\Gamma }_a`$ and $`\gamma _a`$, i.e. we treat $`\mathrm{\Psi }`$ as a 4-component spinor suppressing its extra 4 spectator indices. $$\mathrm{\Gamma }_0=i\sigma _2\times I_8,\mathrm{\Gamma }_1=\sigma _1\times I_8,\tau _3=\mathrm{\Gamma }_0\mathrm{\Gamma }_1=\sigma _3\times I_8,$$ where $`\sigma _{1,2,3}`$ are the Pauli matrices, we end up with 8 species of 2-d Majorana fermions living on a curved 2-d surface with a $`\sigma _3`$ mass term. Assuming that fermions are normalized with $`\sqrt{g}`$, the square of the resulting fermionic operator is then ($`\tau _3^{}=\tau _3`$)<sup>26</sup> Note that Dirac operator is self-adjoint with the measure $`\sqrt{g}`$. $$\mathrm{\Delta }_F^{}=(D_F)^2=(i\tau ^i\widehat{}_i\tau _3)(i\tau ^j\widehat{}_j\tau _3)=\widehat{}^2+\frac{1}{4}R^{(2)}+1,$$ where $`\widehat{}^2=\frac{1}{\sqrt{g}}\widehat{}^i(\sqrt{g}\widehat{}_i).`$<sup>27</sup> Since $`\tau _i=e_i^\alpha \mathrm{\Gamma }_\alpha `$, where $`\mathrm{\Gamma }_\alpha `$ are constant, and since $`\widehat{}`$ is the covariant 2-d spinor derivative, the squaring relation is exactly the same as for the standard 2-d fermions in curved space. Explicitly (recall that here we use the Minkowski signature) $$\begin{array}{cc}\hfill \mathrm{\Delta }_F^{}& =\left[y^1\left(\mathrm{\Gamma }^0_0+\frac{1}{2}y^3y^{}\mathrm{\Gamma }_1\right)+y^3\mathrm{\Gamma }^1_1\right]^2+1\hfill \\ & =y^2\left(_0+\frac{1}{2}y^3y^{}\mathrm{\Gamma }_0\mathrm{\Gamma }_1\right)^2y^4_1\left(y^2_1\right)+\frac{1}{2}\left(1y^4\right).\hfill \end{array}$$ A similarly looking result for the fermionic operator was found in where a different $`\kappa `$-symmetry gauge was used.<sup>28</sup> The kinetic part of our operator is actually different from the expression in which contained an additional connection term in the covariant derivative, and our derivation of the action is much more straightforward. After submission of this paper S. Förste pointed out to us that the two expressions might be related to each other by a rotation, i.e. are equivalent. One can also consider the quadratic fermionic action (6.1) in the “3-brane” gauge $`\theta ^1=i\mathrm{\Gamma }_4\theta ^2`$ (5.1). It was found in that the sum of the quadratic fermionic terms in the action of takes the simple form $$\begin{array}{cc}\hfill S_{2\mathrm{F}}=& 2id^2\sigma \left(\sqrt{g}g^{ij}y^2\overline{\vartheta }\mathrm{\Gamma }_i_j\vartheta ϵ^{ij}_iy\overline{\vartheta }\mathrm{\Gamma }^4_j\vartheta \right)\hfill \\ \hfill =& 2id^2\sigma (y^{1/2}\overline{\theta })\left[\mathrm{\Gamma }^1_1+(y^4\mathrm{\Gamma }^0+y^{}\mathrm{\Gamma }^4)_0\right](y^{1/2}\theta ),\hfill \end{array}$$ where $`\theta \theta ^1`$ is the original GS target space spinor variable related to the rescaled field $`\vartheta `$ in by $`\theta =y^{1/2}\vartheta `$. Here $`\mathrm{\Gamma }^a`$ are constant Dirac matrices ($`\mathrm{\Gamma }^0=\mathrm{\Gamma }_0`$, $`\mathrm{\Gamma }^m=\mathrm{\Gamma }_m`$, $`m=1,4`$). As was noted in , the resulting fermionic operator is non-degenerate. At first sight, this operator is very different from the one in (6.1); in particular, it has no mass term. But the two are, in fact, closely related! To demonstrate this let us note that the combination of the $`\mathrm{\Gamma }`$-matrices multiplying $`_0`$ is actually a local Lorentz rotation of $`\mathrm{\Gamma }^0`$ in the 0-4 plane with parameter $`\beta `$ $$𝒮=\mathrm{exp}\left(\frac{\beta }{2}\mathrm{\Gamma }^0\mathrm{\Gamma }^4\right),\mathrm{cosh}\beta =y^2,\mathrm{sinh}\beta =y^2y^{},\beta ^{}=2y^3,$$ i.e. $$𝒮\mathrm{\Gamma }^0𝒮^1=\mathrm{cosh}\beta \mathrm{\Gamma }^0+\mathrm{sinh}\beta \mathrm{\Gamma }^4=y^2\mathrm{\Gamma }^0+y^2y^{}\mathrm{\Gamma }^4.$$ Introducing $$\chi =𝒮^1\theta ,$$ we get (note that $`\overline{\theta }=\overline{\chi }𝒮^1`$ for any $`SO(9,1)`$ rotation) $$L_{2\mathrm{F}}=2i\overline{\chi }\left[y\mathrm{\Gamma }^0_0+y^1\mathrm{\Gamma }^1_1\frac{1}{2}y^2y^{}\mathrm{\Gamma }_1+y^2\mathrm{\Gamma }^0\mathrm{\Gamma }^1\mathrm{\Gamma }^4\right]\chi .$$ To put this action into the ‘curved space 2-d spinor’ form we need to make a redefinition $`\chi \psi `$ $$\chi =y\psi ,$$ $$L_{2\mathrm{F}}=2i\overline{\psi }\left[y^3\mathrm{\Gamma }^0\left(_0+\frac{1}{2}y^3y^{}\mathrm{\Gamma }_0\mathrm{\Gamma }_1\right)+y\mathrm{\Gamma }^1_1+y^4\mathrm{\Gamma }^0\mathrm{\Gamma }^1\mathrm{\Gamma }^4\right]\psi .$$ Since $`i\mathrm{\Gamma }_4=\mathrm{\Gamma }_0\mathrm{\Gamma }_1\mathrm{\Gamma }_2\mathrm{\Gamma }_3`$, i.e. $$\mathrm{\Gamma }^0\mathrm{\Gamma }^1\mathrm{\Gamma }^4=i\mathrm{\Gamma }^2\mathrm{\Gamma }^3,$$ we finally get, using (6.1),(6.1), the expression that is essentially equivalent to (6.1) $$L_{2\mathrm{F}}=2i\sqrt{g}\left(\overline{\psi }\tau ^i\widehat{}_i\psi +i\overline{\psi }\tau _3^{}\psi \right),\tau _3^{}\mathrm{\Gamma }_2\mathrm{\Gamma }_3,\left(\tau _3^{}\right)^2=1.$$ The square of the fermionic operator in (6.1) is indeed the same as in (6.1): $$\mathrm{\Delta }_F^{}=(D_F)^{}(D_F)=\left(i\tau ^i\widehat{}_i+\tau _3^{}\right)\left(i\tau ^j\widehat{}_j\tau _3^{}\right)=\widehat{}^2+\frac{1}{4}R^{(2)}+1,$$ where we used that, in contrast to $`\tau _3`$ in (6.1), $`\tau _3^{}`$ is antihermitean and commutes with $`\tau _i`$. Since $`\theta `$ and thus its image under the rotation $`\chi `$ are assumed to have canonical normalization, $`\chi ^2=d^2\sigma \sqrt{g}\overline{\chi }\chi `$, we conclude that $`\psi `$ in (6.1) should be normalized with the extra factor of $`y^2`$. As in the case of the circle in Section 5.1, this extra factor is offset by the non-trivial $`\kappa `$-symmetry ghost determinant (5.1) corresponding to the gauge $`\theta ^1=i\mathrm{\Gamma }_4\theta ^2`$, so the fermionic contributions in the two $`\kappa `$-symmetry gauges are again equivalent. 6.4. Partition function Let us first combine the bosonic contributions. In the conformal gauge $$Z_{\mathrm{bose},\mathrm{conf}.\mathrm{g}.}=\frac{\stackrel{1/2}{det}\left(_{ij}^2\frac{1}{2}R^{(2)}g_{ij}\right)}{det^{1/2}\left(D_{ab}^2+X_{ab}\right)det^{5/2}\left(^2\right)},$$ while in the static gauge $$Z_{\mathrm{bose},\mathrm{stat}.\mathrm{g}.}=\frac{1}{det^{2/2}\left(^2+2\right)det^{1/2}\left(^2+R^{(2)}+4\right)det^{5/2}\left(^2\right)}.$$ Up to global factors in the gauge group, the two expressions must be equivalent; it is easy to see, for example, that the corresponding logarithmic divergence coefficients are indeed the same $$(b_2)_{\mathrm{bose},\mathrm{conf}.\mathrm{g}.}=(5+52)\times \frac{1}{6}R^{(2)}8R^{(2)},$$ $$(b_2)_{\mathrm{bose},\mathrm{stat}.\mathrm{g}.}=(2+1+5)\times \frac{1}{6}R^{(2)}2\times 24R^{(2)},$$ where we used (6.1). The contributions of the massless determinants and the ghost determinant can be found, as usual, by integrating the conformal anomaly (see Appendix A), but to find the massive determinants one needs to solve the corresponding spectral problems. Including fermions, the expression for the 1-loop partition function of a string in $`AdS_5\times S^5`$ background with world surface ending on two parallel lines is thus $$Z_{AdS_5\times S^5}^{}=\frac{\stackrel{8/2}{det}(\widehat{}^2+\frac{1}{4}R^{(2)}+1)}{det^{2/2}\left(^2+2\right)det^{1/2}\left(^2+R^{(2)}+4\right)det^{5/2}\left(^2\right)},$$ which is essentially the same as in . Here we took the fermionic contribution in the $`\theta ^1=\theta ^2`$ gauge, and the bosonic contribution in the static gauge. The geometry under consideration is asymptotic to $`AdS_2`$. For example, if we change the coordinate $`\sigma `$ to $`y`$ the induced metric takes the form<sup>29</sup> It is sometimes useful to use the coordinate $`w=1/y`$, in terms of which the metric is $`ds^2=\frac{1}{w^2}[d\tau ^2+\frac{w_0^4}{w_0^4w^4}dw^2]`$, where $`0<w<w_0=1/y_0`$ and $`w=0`$ corresponds to the boundary. $$ds^2=y^2d\tau ^2+\frac{y^2}{y^4y_0^4}dy^2,y_0y<\mathrm{}.$$ The $`y_0=0`$ limit of (6.1) corresponds to the straight string configuration where the metric becomes that of Euclidean $`AdS_2`$ space (with $`0y<\mathrm{}`$). As usual in the case of negative curvature non-compact spaces similar to $`AdS`$, the corresponding actions will in general be divergent and one will need to add boundary counterterms. The details of how the divergent boundary behavior is properly accounted for is actually rather irrelevant for the purpose of extracting the non-trivial finite $`\frac{T}{L}`$ part of $`\mathrm{ln}Z_{AdS_5\times S^5}^{}`$ we are interested in. To avoid altogether questions about boundary terms (and details of topological infinity cancellation) we may normalize our partition function for each field by the partition function of an equivalent field in the straight string configuration, i.e. divide the partition function (6.1) for the noncompact hyperbolic negative curvature space (6.1) by the partition function (twice, to account for the two asymptotic regions), (4.1) for the $`AdS_2`$ case $$\overline{Z}_{AdS_5\times S^5}^{}(y_0)=\frac{Z_{AdS_5\times S^5}^{}(y_0)}{Z_{AdS_5\times S^5}^{}(0)}.$$ Since the topology and the near-boundary (large $`y`$) behavior of the two metrics is the same, this eliminates the problem of carefully tracking down all boundary terms in the expressions for the determinants and allows us to ignore the boundary contributions as well as the total derivative bulk terms (such as the logarithmically divergent terms proportional to $`d^2\sigma \sqrt{g}R^{(2)}`$).<sup>30</sup> It is easy to see that the divergent integral $`d^2\sigma \sqrt{g}R^{(2)}`$ gets contribution only from the boundary behavior of the metric. The ratio of the determinants for the metric (6.1) and for its $`y_0=0`$ limit will be finite and well-defined. This is actually the standard recipe of defining the determinants of Laplace operators on (e.g. 2-dimensional) non-compact spaces by using fiducial metrics of constant negative curvature which have the same topology and asymptotic behavior. As a result, one will need to compute only the well-defined heat kernels like $`\mathrm{Tr}[e^{t\mathrm{\Delta }(y_0)}e^{t\mathrm{\Delta }(0)}]`$. ¿From a practical point of view, the subtraction of the $`AdS_2`$ contributions allows us, as in the evaluation of the classical action (6.1),(6.1), to freely integrate by parts and to drop all divergent boundary contributions. Some examples illustrating this procedure are given in Appendix A.3. 6.5. Crude approximation for the 1-loop potential One may make the following very simple (but probably too crude) estimate of the value of the resulting partition function and thus of the coefficient in the 1-loop correction $`c_1/L`$ to the potential. The classical solution $`y`$ as a function of $`\sigma `$ is approximately equal to $`y_0`$ and changes slowly near $`\sigma 0`$ and then blows up to infinity at the boundaries of the $`\sigma `$-interval $`(L/2,L/2)`$. It seems reasonable to assume that the near-boundary behavior of $`y(\sigma )`$ should not be very important for the value of the normalized partition function (6.1). One may then approximate $`y(\sigma )`$ to be made of three straight pieces. A part where $`yy_0,y^{}0,y^{\prime \prime }0`$, and two parts connecting it to the boundary. Note that this is not the same as taking the flat space limit, since now in (6.1) we have determinants of operators with non-zero mass. Below we will estimate the contribution of the part at $`yy_0`$. We did not evaluate the contribution from the two pieces connecting it to the boundary. Since $`y`$ is assumed to change very slowly, we may set $`R^{(2)}0`$. Then we are left with the following combination of determinants in flat metric<sup>31</sup> Contributions of overall constant factors like $`y_0^2`$ in the operators cancel out due to supersymmetric balance of the numbers of fields. $$\begin{array}{cc}\hfill W=& \frac{1}{2}[2\mathrm{ln}det(^2+2y_0^2)+\mathrm{ln}det(^2+4y_0^2)+5\mathrm{ln}det(^2)\hfill \\ & 8\mathrm{ln}det(^2+y_0^2)].\hfill \end{array}$$ This effective action is UV finite because of the obvious mass sum rule. The non-zero finite result for $`W`$ can probably be interpreted as the vacuum energy of some spontaneously broken supersymmetric 2-d field theory corresponding to (6.1). Assuming Dirichlet boundary conditions in both $`\tau `$ and $`\sigma `$ directions and taking $`T\mathrm{}`$ (so that we can integrate over the continuous eigenvalue in $`\tau `$-direction) we get<sup>32</sup> In general, for a massive determinant we get $`_{n,k=1}^{\mathrm{}}\left[(\frac{\pi n}{T})^2+(\frac{\pi k}{L})^2+m^2\right]`$. For a massless determinant $`_{n,m=1}^{\mathrm{}}\left[(\frac{\pi n}{T})^2+(\frac{\pi m}{L})^2\right]=(2L)^{1/2}\eta (i\frac{T}{L}),`$ where $`\eta (x)=e^{i\frac{\pi }{12}x}_{n=1}^{\mathrm{}}(1e^{i\pi nx})`$. $$\begin{array}{cc}& \frac{1}{2}\mathrm{ln}det(^2+m^2)\frac{1}{2}\mathrm{ln}det(^2)\hfill \\ & =\frac{1}{2}T\underset{n=1}{\overset{\mathrm{}}{}}_{\mathrm{}}^{\mathrm{}}\frac{dk}{2\pi }\mathrm{ln}\left(1+\frac{\omega _n^2}{k^2}\right)=\frac{1}{2}T\underset{n=0}{\overset{\mathrm{}}{}}\omega _n,\omega _n^2=(\frac{\pi n}{L})^2+m^2.\hfill \end{array}$$ Thus $$W=z_1\frac{\pi T}{2L},z_1=\underset{n=1}{\overset{\mathrm{}}{}}\left[\sqrt{n^2+4a^2}+2\sqrt{n^2+2a^2}+5\sqrt{n^2}8\sqrt{n^2+a^2}\right],$$ $$a=\frac{Ly_0}{\pi }=\frac{2\sqrt{2\pi }}{[\mathrm{\Gamma }(\frac{1}{4})]^2}0.38138,$$ where we have used (6.1). To compare, in the “massless” case one finds \[23\] (see (2.1)) $`z_1=_{n=1}^{\mathrm{}}n=\zeta (1)=\frac{1}{12}0.08333`$. The infinite sum (6.1) is convergent and its the numerical evaluation gives $$z_11.24966.$$ Thus the coefficient of the $`1/L`$ potential is negative, i.e. has the same sign as a boson in flat space. To this one has to add the contribution of the two “flat” lines connecting it to the boundary. This will give a result that is not identical to that of $`AdS_2`$, because these lines extend only up to $`y_0`$. To go beyond the above crude approximation and to compute $`\mathrm{ln}\overline{Z}_{AdS_5\times S^5}^{}=c_1\frac{T}{L}`$ in (6.1) exactly one may use the following strategy: (i) first, one may compute the contributions of the massless determinants and the Jacobian using the results of Appendices A and C; (ii) then, since the induced 2-d metric is conformally flat, one may rescale it to the flat space one, isolating the conformal anomaly parts of the determinants; (iii) finally, one may compute the spectra of the resulting operators in flat metric with $`y`$-dependence being only in the mass terms. For example, the bosonic operators in (6.1) then become $$_0^2_1^2+2y^2,_0^2_1^2+2y^22y^2,_0^2_1^2,$$ where we have made the coordinate change $`\sigma \sigma ^{}`$ such that the 2-d metric becomes conformally flat, $$ds^2=y^2(d\tau ^2+d\sigma ^2),d\sigma ^{}=y^2d\sigma .$$ The computation of the spectra of these operators defined in the 2-d strip $`(T,L^{})`$, where $`L^{}=\frac{[\mathrm{\Gamma }(\frac{1}{4})]^4}{2(2\pi )^2}L`$ is the range of $`\sigma ^{}`$ (see Appendix A.3), is left for the future. 7. Conclusions We have presented a systematic treatment of the Green Schwarz string in curved $`AdS_5\times S^5`$ space. We found the quadratic fluctuation operators in conformal and static gauges (for Polyakov and Nambu-Goto actions). A careful treatment was presented of the measure factors and ghost determinants. We also considered two different ways of gauge fixing the $`\kappa `$-symmetry, and explained how one can relate the GS fermion kinetic term to the standard 2-d Dirac fermion action on a curved 2-d background by making a local target-space Lorentz rotation. We discussed the resolution of the problem of the logarithmic $`O(R^{(2)})`$ divergences encountered (in the Nambu framework) in . First, as in the case of the flat target space, the divergence proportional to $`R^{(2)}`$ should be accompanied by a boundary term, promoting it to the Euler number, and thus is topological. The cancellation of topological divergences in a critical string theory is implied by careful definition of the path integral measure. This is clearer in the Polyakov approach, but should also be true in the Nambu formulation. In any case, this issue does not arise in the case of the induced 2-d geometries that we discussed (except for the circle), since there the Euler number vanishes. We have emphasized that the natural way to define the partition function in the case where the induced geometry is asymptotic to $`AdS_2`$ (the case relevant for computing the correction to quark – anti-quark potential) is to normalize with respect to the partition function for $`AdS_2`$ space. Then the issues of boundary counterterms and divergences simply do not arise. We have studied the cases of minimal surfaces ending along a straight line, a circle and two lines. In the first two cases we ended up with a supersymmetric field theory on $`AdS_2`$. The straight line is a BPS object, and therefore one expects $`Z=1`$, as we were able to verify. We presented two other ways of calculating the partition function on $`AdS_2`$, which give different results. The discrepancy is attributed to different regularizations and to assumptions about the asymptotics of the eigenfunctions. Those calculations might be more appropriate for the circular loop geometry, where supersymmetry is broken. In the case of the two parallel lines we have found the general expression for the partition function and showed how to express it in terms of the determinants of 2-d Laplace operators on a flat strip with potentials depending only on one of the two coordinates. We have not, however, addressed the issue of finding exact analytical or numerical methods of computing the corresponding determinants and thus the value of the numerical constant in the subleading correction to the $`1/L`$ potential. Acknowledgements N.D. would like to thank G. Horowitz, N. Itzhaki and J. Polchinski, and A.A.T. would like to thank R. Metsaev, P. Olesen and A. Polyakov for useful discussions of related questions. The work of N.D. and D.J.G. is supported by the NSF under grant No. PHY94-07194. The work of A.T. is supported by the DOE grant DOE/ER/01545-780, EC TMR grant ERBFMRX-CT96-0045, INTAS grant No.96-538 and NATO grant PST.CLG 974965. Part of this work was done while A.A.T. was visiting the Institute of Theoretical Physics at UCSB, and he would like to acknowledge the hospitality of ITP extended to him and the support of the NSF grant No. PHY94-07194. Appendix A. The dependence of determinants on measure and conformal factors A.1. Bosonic operators We start with a review of some general facts about divergences and conformal anomalies of 2-d determinants. In particular, we review the issue of the measure dependence of the determinants \[49,,50\]. The action $$\begin{array}{cc}& S_2=\frac{1}{2}d^2\sigma \sqrt{g}(g^{ij}_i\varphi _j\varphi +X\varphi ^2)(\varphi ,\mathrm{\Delta }\varphi ),\hfill \\ & (\varphi ,\varphi ^{})d^2\sigma \sqrt{g}M(\sigma )\varphi (\sigma )\varphi ^{}(\sigma ),\hfill \end{array}$$ where the scalar product is defined with an extra measure factor $`M`$, implies that the relevant Laplace operator that occurs in the determinant is, not $`\mathrm{\Delta }`$, but rather $$\mathrm{\Delta }_M=M^1(^2+X).$$ The dependence of $`det\mathrm{\Delta }_M`$ on $`M`$ can be determined by using the standard observation that, since $`\delta \mathrm{\Delta }_M=(M^1\delta M)\mathrm{\Delta }_M,`$ the variation of $$\mathrm{ln}det\mathrm{\Delta }_M=_{\mathrm{\Lambda }^2}^{\mathrm{}}\frac{dt}{t}\mathrm{Tr}\mathrm{exp}(t\mathrm{\Delta }_M)$$ with respect to $`\mathrm{ln}M`$ can be expressed in terms of the Seeley coefficients of $`\mathrm{\Delta }_M`$. It is then easy to see that only the quadratically (and linearly) divergent and finite parts of $`\mathrm{ln}det\mathrm{\Delta }_M`$ are dependent on $`M`$, but that the logarithmically divergent part is $`M`$-independent. This is in agreement with naive expectation that the $`M`$ dependence should be given by $`\mathrm{ln}M`$ multiplying a regularized “$`\delta (0)`$”. Equivalently, note that (A.1) can be written as $$\mathrm{\Delta }_M=\stackrel{~}{}^2+\stackrel{~}{X},\stackrel{~}{g}_{ab}Mg_{ab},\stackrel{~}{X}M^1X.$$ Then the divergent part of this determinant is given by the standard expression $$\begin{array}{cc}\hfill (\mathrm{ln}det\mathrm{\Delta }_M)_{\mathrm{}}=& \frac{1}{4\pi }\mathrm{\Lambda }^2d^2\sigma \sqrt{\stackrel{~}{g}}\pm \frac{1}{4\sqrt{\pi }}\mathrm{\Lambda }𝑑s\sqrt{\stackrel{~}{\gamma }}\hfill \\ & \frac{1}{4\pi }\mathrm{ln}\mathrm{\Lambda }^2\left[d^2\sigma \sqrt{\stackrel{~}{g}}\left(\frac{1}{6}\stackrel{~}{R}\stackrel{~}{X}\right)+\frac{1}{3}𝑑s\sqrt{\stackrel{~}{\gamma }}\stackrel{~}{K}\right],\hfill \end{array}$$ where $`\pm `$ corresponds to the Dirichlet and Neumann boundary conditions, $`\stackrel{~}{\gamma }=M\gamma `$ is the boundary metric and $`K`$ is the trace of the second fundamental form. It is easy to see that the dependence on $`M`$ drops out of the $`\mathrm{ln}\mathrm{\Lambda }`$ term.<sup>33</sup> This is not unexpected since the relevant part of the logarithmic divergence is proportional to the Euler number of the $`\stackrel{~}{g}`$ metric but it should be the same as the Euler number of the $`g`$ metric. Strictly speaking, this is so if $`M`$ is smooth inside the domain, so as not to change the topology. The dependence of the finite part of $`\mathrm{ln}det\mathrm{\Delta }_M`$ on $`M`$ is dictated by the same Seeley coefficient $`a_2`$ which multiples the $`\mathrm{ln}\mathrm{\Lambda }`$ term and determines the conformal anomaly. This coefficient, $`a_2`$, appears in the $`t0`$ expansion of the heat kernel, $$\mathrm{Tr}[F(\sigma )\mathrm{exp}(t\mathrm{\Delta }_M)]=\frac{1}{t}a_0(F|\mathrm{\Delta }_M)+\frac{1}{\sqrt{t}}a_1(F|\mathrm{\Delta }_M)+a_2(F|\mathrm{\Delta }_M)+O\left(\sqrt{t}\right),$$ where $`F`$ is an arbitrary function, and $$\begin{array}{cc}\hfill a_0(F|\mathrm{\Delta }_M)=& \frac{1}{4\pi }d^2\sigma \sqrt{g}M(\sigma )F(\sigma ),\hfill \\ \hfill a_1(F|\mathrm{\Delta }_M)=& \frac{1}{8\sqrt{\pi }}𝑑s\sqrt{\gamma }\sqrt{M(s)}F(s),\hfill \\ \hfill a_2(F|\mathrm{\Delta }_M)=& \frac{1}{4\pi }\left[d^2\sigma \sqrt{g}F(\sigma )b_2(\mathrm{\Delta }_M)+𝑑s\sqrt{\gamma }\left(F(s)c_2(\mathrm{\Delta }_M)\frac{1}{2}\sqrt{\gamma }_nF\right)\right].\hfill \end{array}$$ Here $$\begin{array}{cc}\hfill b_2(\mathrm{\Delta }_M)=& \frac{1}{6}R^{(2)}X\frac{1}{6}^2\mathrm{ln}M,\hfill \\ \hfill c_2(\mathrm{\Delta }_M)=& \frac{1}{3}\left(K\frac{1}{2}_n\mathrm{ln}M\right).\hfill \end{array}$$ and $`_n`$ is the inward pointing derivative normal to the boundary. The signs $``$ are for D and N boundary conditions. The dependence of the finite part on the measure factor $`M`$ is found by integrating the equation (we assume that $`\mathrm{\Delta }_M`$ has a trivial kernel) $$\delta (\mathrm{ln}det\mathrm{\Delta }_M)=a_2(\delta \mathrm{ln}M|\mathrm{\Delta }_M),$$ i.e. $$\begin{array}{cc}\hfill (\mathrm{ln}det\mathrm{\Delta }_M)_{\mathrm{fin}}=& (\mathrm{ln}det\mathrm{\Delta }_1)_{\mathrm{fin}}\frac{1}{4\pi }d^2\sigma \sqrt{g}\left[\mathrm{ln}M\left(\frac{1}{6}R^{(2)}X\right)+\frac{1}{12}^i\mathrm{ln}M_i\mathrm{ln}M\right]\hfill \\ & \frac{1}{12\pi }𝑑s\sqrt{\gamma }\left(\mathrm{ln}MK\frac{1}{2}_n\mathrm{ln}M\right).\hfill \end{array}$$ A.2. Fermionic operators Consider now the fermionic action $$d^2\sigma \sqrt{g}\overline{\psi }e_\alpha ^i\tau ^\alpha D_i\psi =d^2\sigma \sqrt{g}\overline{\psi }D_F\psi ,$$ where $`\tau _\alpha `$ are 2-d Dirac matrices. Assume that the norm contains an extra function $`𝒦`$ $$\psi ^2=d^2\sigma \sqrt{g}𝒦\overline{\psi }\psi .$$ Then the relevant second order operator is $$\mathrm{\Delta }_𝒦^{(F)}=(𝒦^1D_F)^2=𝒦^2(\widehat{}^2+\mathrm{}).$$ While $`\mathrm{\Delta }_𝒦^{(F)}`$ looks like the Laplace operator (A.1), there are two important differences compared to the scalar case: (i) the fermionic measure is $`𝒦`$, not $`M=𝒦^2`$, and (ii) in addition to the overall factor $`𝒦^2`$, there is also an extra first derivative term, leading to an extra connection and extra potential terms. The dependence on $`𝒦`$ can be found using again the variational argument, as in the derivation of conformal anomaly. The logarithmic divergences again do not depend on $`𝒦`$. The variation of $`(𝒦^1D_F)^2`$ over $`𝒦`$ is the same as of $`𝒦^2(D_F)^2`$, but now it is the Seeley coefficient of $`(𝒦^1D_F)^2`$ that is to be used in (A.1). In more detail, set $`𝒦=e^\lambda `$ and choose the conformal frame $`e_i^\alpha =e^\rho \delta _i^\alpha `$. Then it is easy to show that the since the spinor derivative is $`D_j=_j+\frac{1}{2}i\tau _3ϵ_{jk}_k\rho `$, where the index contractions are with respect to the flat metric, the operator $`(𝒦^1D_F)^2`$ becomes (we add here a potential term $`Y`$ for generality) $$\begin{array}{cc}\hfill \mathrm{\Delta }_𝒦^{(F)}=& \left(e^\lambda e^{\frac{3}{2}\rho }\tau _i_ie^{\frac{1}{2}\rho }\right)^2+e^{2\lambda }Y\hfill \\ \hfill =& e^{2\lambda 2\rho }\left(\delta ^{ij}+iϵ^{ij}\tau _3\right)_i(1\lambda \frac{1}{2}\rho )_j(1+\frac{1}{2}\rho )+e^{2\lambda }Y,\hfill \end{array}$$ where $`_i`$ and $`_j`$ act on all terms to the right. This can be put into the standard form (A.1) (with $`M=𝒦^2`$) as follows $$\begin{array}{cc}\hfill \mathrm{\Delta }_𝒦^{(F)}=& e^{2\lambda }\left[e^{2\rho }(_k+B_k)^2+X\right],\hfill \\ \hfill _k+B_k=& _k+\frac{i}{2}ϵ_{kj}\tau _3(_j\rho +_j\lambda )\frac{1}{2}_k\lambda =D_k\frac{1}{2}\tau _j\tau _k_j\lambda ,\hfill \\ \hfill X=& Y\frac{1}{2}e^{2\rho }^2\rho \frac{1}{2}e^{2\rho }^2\lambda =Y+\frac{1}{4}R^{(2)}\frac{1}{2}^2\lambda ,\hfill \end{array}$$ where we have used that in $`d=2`$ $`\tau ^i\tau _j\tau _i=0`$. The corresponding Seeley coefficient is thus (cf. (A.1)) $$b_2\left(\mathrm{\Delta }_𝒦^{(F)}\right)=\left(\frac{1}{6}\frac{1}{4}\right)R^{(2)}Y\left(\frac{1}{3}\frac{1}{2}\right)^2\lambda ,$$ or, in conformal coordinates, $$\sqrt{g}b_2\left(\mathrm{\Delta }_𝒦^{(F)}\right)=\left(\frac{1}{3}\frac{1}{2}\right)^2\rho \left(\frac{1}{3}\frac{1}{2}\right)^2\lambda \sqrt{g}Y,$$ so that $`\lambda `$ enters like the conformal factor (this argument determines the conformal anomaly of the Dirac operator since $`\sqrt{g}e_\alpha ^i=e^\rho \delta _\alpha ^i`$). As in (A.1), we have (omitting obvious boundary terms) $$\delta \left(\mathrm{ln}det\mathrm{\Delta }_𝒦^{(F)}\right)=2a_2\left(\delta \mathrm{ln}𝒦|\mathrm{\Delta }_𝒦^{(F)}\right),$$ and thus $$\begin{array}{cc}\hfill \left(\mathrm{ln}det\mathrm{\Delta }_𝒦^{(F)}\right)_{\mathrm{fin}}=& \left(\mathrm{ln}det\mathrm{\Delta }_1^{(F)}\right)_{\mathrm{fin}}\hfill \\ & +\frac{1}{4\pi }d^2\sigma \sqrt{g}\left[\mathrm{ln}𝒦\left(\frac{1}{6}R^{(2)}+2Y\right)+\frac{1}{6}g^{ij}_i\mathrm{ln}𝒦_j\mathrm{ln}𝒦\right].\hfill \end{array}$$ Note that the conformal anomaly of a scalar is twice as of a 2-d fermion, since for a scalar $`Me^{2\rho }`$ while for a fermion $`𝒦e^\rho `$. The anomaly of a GS spinor is 4 times as much as of 2-d fermion since here one needs to take $`𝒦e^{2\rho }`$ on top of the flat space operator.<sup>34</sup> Note that redefining spinors with careful account of measure factors gives equivalent results. For example, the usual 2-d spinor action is, in conformal gauge, $`d^2\sigma \sqrt{g}\overline{\psi }e^{\frac{3}{2}\rho }\tau ^\alpha _\alpha e^{\frac{1}{2}\rho }\psi `$ with the measure $`d^2\sigma \sqrt{g}\overline{\psi }\psi `$. Redefining $`\psi ^{}=e^{\frac{1}{2}\rho }\psi `$ we get $`d^2\sigma \sqrt{g}\overline{\psi }^{}e^{2\rho }\tau ^\alpha _\alpha \psi ^{}`$ with the measure $`d^2\sigma \sqrt{g}e^\rho \overline{\psi }^{}\psi ^{}`$, i.e. $`𝒦=e^\rho `$. That corresponds to the operator $`𝒦^1D_F=e^\rho \tau ^\alpha _\alpha `$ which has the same determinant as $`e^{\frac{3}{2}\rho }\tau ^\alpha _\alpha e^{\frac{1}{2}\rho }`$. It is useful to compare this with the result found by treating $`\mathrm{\Delta }_𝒦^{(F)}`$ as a scalar operator (A.1) with $`M=𝒦^2`$. According to (A.1) we would get (using that $`X\frac{1}{4}R^{(2)}+X`$ in this case) $$\begin{array}{cc}\hfill \left(\mathrm{ln}det\mathrm{\Delta }_{𝒦^2}\right)_{\mathrm{fin}}=& \left(\mathrm{ln}det\mathrm{\Delta }_1\right)_{\mathrm{fin}}\hfill \\ & \frac{1}{4\pi }d^2\sigma \sqrt{g}\left[\mathrm{ln}𝒦\left(\frac{1}{6}R^{(2)}2X\right)+\frac{1}{3}g^{ij}_i\mathrm{ln}𝒦_j\mathrm{ln}𝒦\right].\hfill \end{array}$$ Thus $$\left(\mathrm{ln}det\mathrm{\Delta }_{𝒦^2}\right)_{\mathrm{fin}}\left(\mathrm{ln}det\mathrm{\Delta }_𝒦^{(F)}\right)_{\mathrm{fin}}=\frac{1}{4\pi }d^2\sigma \sqrt{g}\left[\left(\frac{1}{3}+\frac{1}{6}\right)g^{ij}_i\mathrm{ln}𝒦_j\mathrm{ln}𝒦\right].$$ As expected, there is a non-trivial difference for a non-constant $`𝒦`$. A.3. Explicit results for some determinants As was mentioned in the text, one way to calculate the determinants in a curved geometry is to transform to flat metric. Instead of a complicated kinetic term depending on induced metric one then has to deal with a complicated mass term. Let us consider the expression for a scalar in the general “bent” string configuration of Section 6. The determinant consists of two terms: a flat-space determinant and a conformal anomaly part. For a single massless scalar with canonical normalization the conformal anomaly related part of its bulk effective action is $$\begin{array}{cc}\hfill 𝒲& =\frac{1}{2}\mathrm{ln}det(^2)\frac{1}{2}\mathrm{ln}det(^2)\hfill \\ & =\frac{1}{24\pi }\left[\mathrm{ln}\mathrm{\Lambda }d^2\sigma \sqrt{g}R^{(2)}+\frac{1}{4}R^{(2)}(^2)R^{(2)}\right]\hfill \\ & \frac{1}{24\pi }d^2\sigma ^{}_\alpha \rho _\alpha \rho ,\hfill \end{array}$$ where we have chosen the conformal coordinate system where $$ds^2=e^{2\rho }(d\tau ^2+d\sigma ^2),\rho =\mathrm{ln}y,d\sigma ^{}=\frac{y^2}{y_0^2}d\sigma .$$ To evaluate this integral let us note that at the boundary $`y=\mathrm{}`$. As a result, the total derivative and boundary contributions are trivial divergences which can be ‘renormalized away’ by subtracting the expression for the straight line case ($`y_00`$). This allows us to freely integrate by parts and to replace, e.g., $`y^4y_0^4`$. Dropping the total derivative part we thus get $$𝒲=\frac{1}{24\pi }d^2\sigma ^{}_i\rho _i\rho =\frac{1}{24\pi }𝑑\tau 𝑑\sigma y_0^2\frac{y^2}{y^4}\frac{TL}{12\pi }y_0^2=\frac{2\pi ^2}{3[\mathrm{\Gamma }(\frac{1}{4})]^4}\frac{T}{L}.$$ Eq. (A.1) is to be added to the flat space result (2.1) (we assume Dirichlet boundary condition) $$\frac{1}{2}\mathrm{ln}det(^2)=\frac{\pi }{24}\frac{T}{L^{}}.$$ Here $`L^{}`$ is the range of $`\sigma ^{}`$ which is different from $`L`$ by a factor, $$L^{}=𝑑\sigma \frac{y^2}{y_0^2}=\frac{2\sqrt{\pi }\mathrm{\Gamma }(\frac{5}{4})}{\mathrm{\Gamma }(\frac{3}{4})}\frac{1}{y_0}=\frac{[\mathrm{\Gamma }(\frac{1}{4})]^2}{2\sqrt{2\pi }}\frac{1}{y_0}=\frac{[\mathrm{\Gamma }(\frac{1}{4})]^4}{2(2\pi )^2}L,$$ where we used that $`\mathrm{\Gamma }(\frac{3}{4})\mathrm{\Gamma }(\frac{1}{4})=\sqrt{2}\pi `$. Finally, we get the following expression for the massless scalar determinant $$\frac{1}{2}\mathrm{ln}det(^2)=\frac{(\pi 2)\pi ^2}{3[\mathrm{\Gamma }(\frac{1}{4})]^4}\frac{T}{L}.$$ Like the flat-space potential (2.1) and like the tree-level potential (6.1) this expression is negative. Multiplied by 5, this gives the result for the contribution of massless fluctuations in the $`S^5`$ directions to the partition function (6.1). Other determinants should lead to similar contributions. Let us now consider the case of the fermionic determinant using (A.1). One finds that the conformal anomaly for a massless 2-d spinor is 1/2 of that for the scalar (A.1), i.e. $$𝒲_\mathrm{F}=\frac{1}{48\pi }d^2\sigma ^{}_\alpha \rho _\alpha \rho =\frac{\pi ^2}{3[\mathrm{\Gamma }(\frac{1}{4})]^4}\frac{T}{L}.$$ Appendix B. Partition function in straight string case ($`AdS_2`$) Here we describe a direct approach to the calculation of the partition function (4.1) on $`AdS_2`$ which complements the discussion in Section 4.3. B.1. Spectral density and $`\zeta `$-function on Poincaré disc Our starting point will be the spectrum of the operator (4.1) with Dirichlet conditions at the boundary of the Poincaré disc, following . The trace of heat kernel is defined by $$K_B(t;m^2)=\mathrm{Tr}\mathrm{exp}[t(\mathrm{\Delta }+m^2)]=\frac{1}{2\pi ^2}_0^{\mathrm{}}\mathrm{exp}\left[t\lambda (\nu )\right]\mu (\nu )𝑑\nu ,$$ and the zeta function is $$\zeta _B(s;m^2)=\mathrm{Tr}\frac{1}{(\mathrm{\Delta }+m^2)^s}=\frac{1}{2\pi ^2}_0^{\mathrm{}}\frac{\mu (\nu )d\nu }{\lambda (\nu )^s}.$$ The density of states for a scalar Laplacian $`\mathrm{\Delta }+m^2`$ (in our case $`m^2=0,2`$) is $$\mu _B(\nu )=\pi \nu \mathrm{tanh}(\pi \nu ),$$ where the eigenvalues of the Laplacian are $`\lambda =\nu ^2+m^2+\frac{1}{4}.`$ If we plug in this density of states (B.1) into (B.1) we find (dropping the divergence) $$\zeta _B(s;m^2)=\frac{(m^2+\frac{1}{4})^{(1s)}}{4\pi (s1)}\frac{1}{\pi }_0^{\mathrm{}}\frac{\nu d\nu }{(e^{2\pi \nu }+1)\left(\nu ^2+m^2+\frac{1}{4}\right)^s},$$ where we have used that $`\mathrm{tanh}(\pi \nu )=12/[\mathrm{exp}(2\pi \nu )+1]`$. The divergence in the determinant is proportional to $`\zeta _B(0;m^2)`$ $$\zeta _B(0;m^2)=\frac{1}{4\pi }\left(m^2+\frac{1}{4}\right)+\frac{1}{48\pi }=\frac{1}{4\pi }\left(\frac{1}{6}R^{(2)}m^2\right),$$ while the finite part $$[\mathrm{ln}det(\mathrm{\Delta }+m^2)]_{\mathrm{fin}}=\zeta _B^{}(0;m^2)$$ is found, using (B.1), to be $$\begin{array}{cc}\hfill \zeta _B^{}(0;m^2)=& \frac{1}{4\pi }\left(m^2+\frac{1}{4}\right)\left[\mathrm{ln}\left(m^2+\frac{1}{4}\right)1\right]\hfill \\ & +\frac{1}{\pi }_0^{\mathrm{}}\frac{\nu d\nu }{e^{2\pi \nu }+1}\mathrm{ln}\left(\nu ^2+m^2+\frac{1}{4}\right).\hfill \end{array}$$ The coincidence limit of the propagator is (dropping divergences)<sup>35</sup> To do the integral, we note that $`\frac{1}{e^{2\pi \nu }+1}=\frac{1}{e^{2\pi \nu }1}\frac{2}{e^{4\pi \nu }1}`$, and use $`_0^{\mathrm{}}\frac{\nu d\nu }{(e^{2\pi a\nu }1)\left(\nu ^2+x^2\right)}=\frac{1}{2}\left[\mathrm{ln}(ax)\frac{1}{2ax}\psi (ax)\right]`$ and $`2\psi (2x)\psi (x)=\psi \left(x+\frac{1}{2}\right)+2\mathrm{ln}2`$. $$G_B(0)=\zeta _B(1;m^2)=\frac{1}{4\pi }\mathrm{ln}\left(m^2+\frac{1}{4}\right)\frac{1}{\pi }_0^{\mathrm{}}\frac{\nu d\nu }{(e^{2\pi \nu }+1)\left(\nu ^2+m^2+\frac{1}{4}\right)}$$ $$\begin{array}{cc}\hfill =\frac{1}{2\pi }\psi \left(\frac{1}{2}+\sqrt{m^2+\frac{1}{4}}\right),& \end{array}$$ where $`\psi (x)=\frac{d}{dx}\mathrm{ln}\mathrm{\Gamma }(x)`$. This is equal to $`\frac{d}{dm^2}\zeta _B^{}(0;m^2)`$. Since for $`m^2=\frac{1}{4}`$ we can evaluate $`\zeta ^{}(0;\frac{1}{4})`$ explicitly,<sup>36</sup> To show this use $`_0^{\mathrm{}}\frac{xdx}{e^x+1}=\frac{1}{2}\zeta _R(2)`$, where $`\zeta _R`$ is the Riemann zeta function. and $`_0^{\mathrm{}}\frac{xdx}{e^x+1}\mathrm{ln}x=\frac{1}{2}[(\psi (2)+\mathrm{ln}2)\zeta _R(2)+\zeta _R^{}(2)]`$, and the identities $`\psi (2)=1\gamma `$ and $`\zeta _R(2)=\frac{\pi ^2}{6}`$. we can write $$\begin{array}{cc}\hfill \zeta _B^{}(0;m^2)=\frac{1}{24\pi }(1\gamma \mathrm{ln}\pi )+\frac{1}{4\pi ^3}\zeta _R^{}(2)+\frac{1}{2\pi }_{{\scriptscriptstyle \frac{1}{4}}}^{m^2}𝑑x\psi \left(\frac{1}{2}+\sqrt{x+\frac{1}{4}}\right).& \end{array}$$ The density of states of the Laplacian for a Majorana fermion is $$\mu _F(\nu )=\pi \nu \mathrm{coth}(\pi \nu ),$$ so that the $`\zeta `$-function is $$\zeta _F(s;m^2)=\frac{m^{2(1s)}}{4\pi (s1)}+\frac{1}{\pi }_0^{\mathrm{}}\frac{\nu d\nu }{(e^{2\pi \nu }1)\left(\nu ^2+m^2\right)^s}$$ where we used that $`\mathrm{coth}(\pi \nu )=1+2/[\mathrm{exp}(2\pi \nu )1]`$. The divergence in the determinant is proportional to $`\zeta _F(0;m^2)`$ and $$\zeta _F(0;m^2)=\frac{1}{4\pi }m^2\frac{1}{24\pi }=\frac{1}{4\pi }\left(\frac{1}{6}R^{(2)}\frac{1}{4}R^{(2)}m^2\right).$$ This is the standard result for a 2-d fermion. As discussed in detail in the text (and in Appendix C), in the case of GS fermion in the conformal gauge we should effectively multiply the $`R^{(2)}`$ term in $`\zeta _F(0;m^2)`$ by 4, ensuring the eventual cancellation of conformal anomalies and topological infinities. The finite part of the determinant is $$\begin{array}{cc}\hfill \mathrm{ln}det\left(D_F^2\right)_{\mathrm{fin}.}& =\zeta _F^{}(0;m^2)\hfill \\ & =\frac{1}{4\pi }m^2\left(\mathrm{ln}m^21\right)+\frac{1}{\pi }_0^{\mathrm{}}\frac{\nu d\nu }{e^{2\pi \nu }1}\mathrm{ln}\left(\nu ^2+m^2\right).\hfill \end{array}$$ The derivative of $`\zeta `$-function with respect to $`m^2`$ $$\begin{array}{cc}\hfill \frac{d}{dm^2}\zeta _F^{}(0;m^2)& =\frac{1}{4\pi }\mathrm{ln}m^2+\frac{1}{\pi }_0^{\mathrm{}}\frac{\nu d\nu }{\left(e^{2\pi \nu }1\right)\left(\nu ^2+m^2\right)}\hfill \\ & =\frac{1}{4\pi }\left(\frac{1}{|m|}+2\psi (|m|)\right)\hfill \end{array}$$ is different from the fermion propagator from (divided by $`2m`$). The expression (B.1) is obviously independent of the sign of the fermion mass term. However, supersymmetry relates fermions with opposite masses to scalars with different masses, and the supersymmetric regularization used in gave a propagator that does depend on the sign. for the computation of the partition function, The prescription of represents a different regularization scheme and thus leads to a different expression for the effective potential or partition function (see below). Again, since we can calculate the finite part of the partition function explicitly at $`m=0`$, we get $$\zeta _F^{}(0;m^2)=\frac{1}{12\pi }(1\gamma \mathrm{ln}2\pi )\frac{1}{2\pi ^3}\zeta _R^{}(2)+\frac{|m|}{2\pi }+\frac{1}{2\pi }_0^{m^2}𝑑x\psi (\sqrt{x}).$$ An $`𝒩=1`$ supermultiplet in $`AdS_2`$ contains a fermion of mass $`m_F=\mu `$ and a boson of mass squared $`m_B^2=\mu ^2\mu `$ (4.1). One can combine (B.1) and (B.1) to get the partition function of a single multiplet. This can be written in terms of complicated special functions; for $`\mu =\pm 1`$ the numerical results are $`\zeta _B^{}(0;0)\zeta _F^{}(0;1)0.02688`$ and $`\zeta _B^{}(0;2)\zeta _F^{}(0;1)0.05269`$. B.2. “Effective potential” in $`AdS_2`$ An alternative way to compute the partition function is to start with the Green’s functions (defined in a way consistent with supersymmetry) and to integrate them over the mass parameter to obtain the effective potential as in \[36,,35\]. Following for a multiplet of one boson and one fermion with masses related as in (4.1) we get $$V_{\mathrm{eff}}(\mu )=\frac{1}{2}_0^{m_B^2=\mu ^2\mu }𝑑m^2G(m^2)\frac{1}{2}_0^{m_F=\mu }𝑑m\mathrm{\hspace{0.17em}2}mG(m^2m),$$ where we normalized the effective potential to be zero in the case of the massless multiplet ($`\mu =0`$). Here $$G(m^2)G_B(x,x|m^2)=x\left|(^2+m^2)^1\right|x$$ $$=\frac{1}{4\pi }\left[\mathrm{ln}(c\mathrm{\Lambda }^2)+2\psi \left(\frac{1}{2}+\sqrt{\frac{1}{4}+m^2}\right)\right],$$ where $`\mathrm{\Lambda }`$ is UV cutoff. In dimensional regularization ($`d2`$) we get $`\mathrm{ln}(c\mathrm{\Lambda }^2)=\frac{2}{2d}+\mathrm{ln}(4\pi \mathrm{\Lambda }^2a^2)\gamma `$. It is assumed that fermions are treated using dimensional reduction regularization which preserves supersymmetry, so that the fermionic Green function satisfies $`\mathrm{tr}G_F(x,x|\mu )=2\mu G_B(x,x|\mu ^2\mu )`$. Then $$\mathrm{ln}Z=W_0+d^2\sigma V_{\mathrm{eff}}(\mu ),$$ where $`W_0`$ is the contribution of the massless multiplet that can be found by integrating the conformal anomaly. This gives an alternative way to compute the partition function. In our case of 3 multiplets with $`\mu =1`$ and 5 multiplets with $`\mu =1`$ we get a logarithmically divergent result: the constant ($`m`$-independent) part of $`G(m^2)`$ is multiplied by $$3\left[\frac{1}{2}_2^0𝑑m^2+\frac{1}{2}_1^0𝑑m\mathrm{\hspace{0.17em}2}m\right]+5\left[\frac{1}{2}_0^0𝑑m^2+\frac{1}{2}_1^0𝑑m\mathrm{\hspace{0.17em}2}m\right]$$ $$=\frac{1}{2}(35)=1.$$ The finite part is also non-zero. For a single multiplet $$[V_{\mathrm{eff}}(\mu )]_{\mathrm{fin}}=\frac{1}{4\pi }[_{\mu ^2\mu }^0dm^2\psi (\frac{1}{2}+\sqrt{\frac{1}{4}+m^2})$$ $$_\mu ^0dm\mathrm{\hspace{0.17em}2}m\psi (\frac{1}{2}+|m\frac{1}{2}\left|\right)]=\frac{1}{4\pi }^0_\mu dm\psi (\frac{1}{2}+|m\frac{1}{2}|),$$ where we have changed the integration variable in the first term ($`m^2m^2m`$).<sup>37</sup> The observation that for a supermultiplet the two terms combine in this way was made in and implicitly in . Notice that since $`_b^a𝑑m\psi (m)=\mathrm{ln}\mathrm{\Gamma }(a)\mathrm{ln}\mathrm{\Gamma }(b),`$ the resulting integral is easily computable by splitting the interval and changing the variables. For $`\mu =1`$ we get explicitly $$[V_{\mathrm{eff}}(1)]_{\mathrm{fin}}=\frac{1}{4\pi }_1^0𝑑m\psi (1m)=\frac{1}{4\pi }_1^2𝑑m\psi (m)=0.$$ For $`\mu =1`$ we get zero bosonic contribution and thus $$\begin{array}{cc}\hfill [V_{\mathrm{eff}}(1)]_{\mathrm{fin}}=& \frac{1}{4\pi }_0^1𝑑m\mathrm{\hspace{0.17em}2}m\psi \left(\frac{1}{2}+\left|m\frac{1}{2}\right|\right)\hfill \\ \hfill =& \frac{1}{4\pi }_{1/2}^1𝑑m\mathrm{\hspace{0.17em}2}\psi (m)=\frac{1}{4\pi }\mathrm{ln}\pi ,\hfill \end{array}$$ where we split the integral into two parts and changed the variable. For 3+5 multiplets we get a non-zero result (note that the first term is actually zero) $$3[V_{\mathrm{eff}}(1)]_{\mathrm{fin}}+5[V_{\mathrm{eff}}(1)]_{\mathrm{fin}}=\frac{5}{4\pi }\mathrm{ln}\pi .$$ The contribution of the 8 massless multiplets can be found in the conformal frame $`ds^2=w^2(dt^2+dw^2)`$ to be $$\begin{array}{cc}\hfill W_0=& 8\times \left(1+\frac{1}{2}\right)\times \frac{1}{24\pi }d^2\sigma \sqrt{g}g^{ij}_i\rho _j\rho \hfill \\ \hfill =& \frac{1}{2\pi }d^2\sigma \sqrt{g},\hfill \end{array}$$ where $`g^{ww}_w\rho _w\rho =1,\rho =\mathrm{ln}w`$. Appendix C. Comments on 2-d determinants In the main text we have performed (following earlier discussions of the GS string in \[19,,13,,16,,14,,15,,17\]) a local rotation to put the GS action (3.1) in a “2-d fermion in curved 2-d space” form. In general, the resulting Jacobian is non-trivial and is given by a Polyakov-Wiegmann expression which is a $`U(1)`$ WZ action. In the case when 2-d metric is kept independent the account of the contribution of the rotation Jacobian is crucial in order to show that the conformal anomaly of a GS fermion is 4 times the naive anomaly of a 2-d fermion – as needed to cancel the conformal anomaly of GS string in flat space. As was already mentioned in the text, the cancellation of (dilatonic part of) the conformal anomaly in the one-loop approximation we considered is exactly the same as in flat space (the curved space-time background changes the $`O(R^{(2)})`$ conformal anomaly only starting with the 2-loop approximation ). Let us summarize some basic facts about these determinants. They are always defined modulo local counterterms of background fields which are to be chosen consistent with the symmetries that are to be preserved (see ). For generality, we consider the chiral case, the non-chiral one is just a combination of the two chiral ones. The standard 2-d Weyl spinor operator is (we use the Euclidean notation where $`\tau _{1,2}=\sigma _{1,2}`$ are the Pauli matrices; here $`k,n,m=1,2`$) $$D(e)=\frac{1}{2}i(1\tau _3)e_\alpha ^k\tau ^\alpha (_k+i\tau _3w_k),\omega _k^{\alpha \beta }=2ϵ^{\alpha \beta }w_k.$$ Then $$\mathrm{ln}detD(e)=\frac{i}{12\pi }I(iw)=\frac{1}{2\times 96\pi }R^{(2)}\left(^2\right)^1(R^{(2)}4i_kw^k).$$ In the case of an abelian gauge field background in flat space $$I(A)=W(T_A,A)=𝑑T_AT_A^1A,(\sqrt{g}g^{mn}+iϵ^{mn})(_n+A_n)T_A=0.$$ In conformal coordinates $`A_z=_z\lambda `$$`\lambda =\mathrm{ln}T_A`$, so that $`I(A)=_z\lambda (\overline{}_z\lambda A_{\overline{z}}),`$ which is different from usual abelian expression $`\lambda \overline{}\lambda `$ by the local counterterm $`A_zA_{\overline{z}}`$.<sup>38</sup> An equivalent (up to a local counterterm) expression is $$\mathrm{ln}detD=\frac{1}{3}Z(w)+Z(A),Z(B)=\frac{1}{4\pi }ϵ^{mn}_mB_n\left(^2\right)^1[ϵ^{mn}_mB_ni_mB^m].$$ Using the conformal gauge and definitions $`g_{mn}=e^{2\rho }\delta _{mn},w_m=_m\lambda +\frac{1}{2}ϵ_{mn}^n\rho ,A_m=_ma+ϵ_{mn}^nb,`$ we get $$\mathrm{ln}detD(e,A)=\frac{1}{4\pi }d^2\sigma \left[\frac{1}{12}^m\rho \left(_m\rho +\frac{1}{2}i_m\lambda \right)^mb\left(_mb+\frac{1}{2}i_ma\right)\right].$$ The gauge field dependent WZ term in the Dirac operator case is proportional to $`I(A)+I(\overline{A})`$, or $$i\left(A\frac{\overline{}}{}A+\overline{A}\frac{}{\overline{}}\overline{A}\right),$$ or, after adding a local $`A\overline{A}`$ term, $`i(\overline{A}\overline{}A)\frac{1}{\overline{}}(\overline{A}\overline{}A)`$. For a Majorana spinor on a curved background only the first (Polyakov) conformal anomaly term is present in (C.1) (with coefficient which is 1/2 of the scalar one), while the imaginary Lorentz-anomaly term cancels out. The gravitational WZW action $`I`$ is simply $`^kb_kb`$, where $`w^k=iϵ^{kn}_nb`$, up to a local counterterm $`w^kw_k`$. If there is also some internal connection acting on flavor indices of fermions, then under the chiral projector in 2 dimensions it can be formally rotated away (this is clear in conformal coordinates) $$D(e,A)=\frac{1}{2}i(1\tau _3)e_\alpha ^m\tau ^\alpha (_m+i\tau _3w_m+A_m)=T_AD(e)T_A^1,$$ where $`T_A`$ is a local rotation “eliminating” $`A_m`$, and $$\mathrm{ln}detD(e,A)=\frac{iN}{12\pi }I(iw)+\frac{i}{4\pi }I(A).$$ Now, consider a more general case we are actually interested in $$D(\rho ,B)=\frac{1}{2}i(1\rho _3)\rho ^m(_m+B_m),$$ where $`\rho _m=\rho _m(x)`$ is an arbitrary $`2N\times 2N`$ representation of the 2-d Dirac algebra for some metric $`g_{mn}(x)`$, satisfying $$\rho _m\rho _n=g_{mn}+ie_{mn}\rho _3,e^{mn}=\frac{1}{\sqrt{g}}ϵ^{mn}.$$ The condition on the connection $`B_m`$ is $$_m\rho _n+[B_m,\rho _n]\mathrm{\Gamma }_{mn}^k\rho _k=0.$$ Since $`g_{mn}=e_m^\alpha e_n^\beta \delta _{\alpha \beta }`$, there exists a local rotation $`S`$ that transforms $`\rho _m`$ into the constant $`2\times 2`$ Pauli matrices times $`N\times N`$ unit matrix times the zweibein $`e_n^\alpha `$ $$S\rho _nS^1=\tau _\alpha \times Ie_n^\alpha .$$ Then $$\begin{array}{cc}\hfill SD(\rho ,B)S^1=& D(e,A),\hfill \\ \hfill _m+i\tau _3w_m+A_m=& S(_m+B_m)S^1,\hfill \end{array}$$ and hence $$\mathrm{ln}detD(\rho ,B)=\frac{iN}{3\pi }I(iw)+\frac{i}{8\pi }I(B).$$ Note that for $`B_m=iw_m`$ (with extra 2 flavors, i.e. $`I(B)=2I(iw)`$) and $`N=1`$ we get back to (C.1), i.e. $`\frac{i}{12\pi }I(iw)`$. For the flat target space ($`B_m`$ comes from integration by parts as required for self-adjointness of the Dirac operator) $$\rho _n=_nx^a\mathrm{\Gamma }_a,B_m=\frac{1}{2}\rho _3_m\rho _3=\frac{1}{2}\rho _n_m\rho ^n.$$ In the case of the non-chiral Dirac fermions (where there is no problem with a definition of the determinant of $`D(\rho ,B)`$ discussed in ) we get simply $$\mathrm{ln}detD(\rho ,B)=\frac{N}{24\pi }R^{(2)}\left(^2\right)^1R^{(2)}+\frac{i}{8\pi }[I(B)\stackrel{~}{I}(B)],$$ where $`\stackrel{~}{I}`$ is defined by (C.1) with parity-reflected condition. In the abelian case $`I(B)\stackrel{~}{I}(B)=0`$, up to a local counterterm. Notice that the conformal anomaly term (here we discuss the Dirac spinor, so that it is twice that of a Majorana GS fermion) is 4 times bigger than for the usual 2-d Dirac fermion. Appendix D. Superstring partition function in $`AdS_3\times S^3`$ with RR flux The same calculation can be done in the case of a superstring in the $`AdS_3\times S^3\times T^4`$ with RR background. The corresponding GS action has the form very similar to $`AdS_5\times S^5`$ one and was discussed in . Since all the classical solutions discussed above depend on only three of the $`AdS_5`$ coordinates, they can be directly embedded in $`AdS_3`$ and are still minimal surfaces. There are some differences in treating the quadratic fluctuations in $`AdS_3\times S^3\times T^4`$: (1) there are no massive fluctuations in the $`x^2,x^3`$ directions which are replaced by extra two massless fluctuations in the toroidal $`T^4`$ directions. (2) only half of the 8 effective 2-d fermions get $`\sigma _3`$ mass term in (4.1), (5.1) and (6.1). In the present case it is natural to split the $`\mathrm{\Gamma }`$-matrices in $`(3+3)+4=6+4`$ way, $`\mathrm{\Gamma }_a=\gamma _a\times I_4`$$`a=0,1,\mathrm{},5`$, where $`\gamma _a`$ are 6-d $`8\times 8`$ matrices. The ‘mass term’ in the covariant derivative (3.1) and in the string action now originates from the sum of the electric and magnetic RR 3-form field strengths, $`x^\mu x^\nu \overline{\theta }\mathrm{\Gamma }_\nu \mathrm{\Gamma }^{abc}\mathrm{\Gamma }_\mu (F_{abc}+F_{abc}^{})`$. The combination of the field strengths produces the $`(1+\mathrm{\Gamma }_7)`$ projection operator, so that only half of the 6-d fermions get a mass term. For example, in the case of the straight line the minimal surface is still $`AdS_2`$, but the set of fluctuation fields is different. There are 7 massless bosons, one mass 2 boson, four massless fermions and four mass 1 fermions. They form one $`𝒩=4`$ multiplet (the dimensional reduction of the $`𝒩=2`$ vector multiplet in $`D=4`$) with three massless and one massive boson and four massive fermions. There are 4 other massless bosons and four massless fermions which can also combine into $`𝒩=4`$ multiplet. One can also check that the vacuum energy as defined by the zeta function again vanishes for this combination of fields. The same conformal anomaly calculation as in section 3.4 gives that $`Z=1`$. In the case of the general bent string configuration (parallel Wilson lines), the analog of the $`AdS_5\times S^5`$ partition function (6.1) takes the form (in static gauge) $$Z_{AdS_3\times S^3\times T^4}^{}=\frac{\stackrel{4/2}{det}\left(\widehat{}^2+\frac{1}{4}R^{(2)}+1\right)\stackrel{4/2}{det}\left(\widehat{}^2+\frac{1}{4}R^{(2)}\right)}{det^{1/2}\left(^2+R^{(2)}+4\right)det^{7/2}\left(^2\right)}.$$ References relax J. Maldacena, “The large-N limit of superconformal field theories and supergravity,” Adv. Theor. Math. Phys. 2, 231 (1998), hep-th/9711200. relax S.S. Gubser, I.R. Klebanov and A.M. Polyakov, “Gauge theory correlators from non-critical string theory,” Phys. Lett. B428, 105 (1998), hep-th/9802109. relax E. Witten, “Anti-de Sitter space and holography,” Adv. Theor. Math. Phys. 2, 253 (1998), hep-th/9802150. relax J. Maldacena, “Wilson loops in large N field theories,” Phys. Rev. Lett. 80, 4859 (1998), hep-th/9803002; S.-J. Rey and J. Yee, “Macroscopic Strings as Heavy Quarks of Large N Gauge Theory and Anti-de Sitter Supergravity,” hep-th/9803001. relax R.R. Metsaev and A.A. Tseytlin, “Type IIB superstring action in $`AdS_5\times S^5`$ background,” Nucl. Phys. B533, 109 (1998), hep-th/9805028. relax R. Kallosh and A.A. Tseytlin, “Simplifying superstring action on $`AdS_5\times S^5`$,” JHEP 10, 016 (1998), hep-th/9808088. relax S. Förste, D. Ghoshal and S. Theisen, “Stringy corrections to the Wilson loop in N = 4 super Yang-Mills theory,” JHEP 08, 013 (1999), hep-th/9903042 relax J. Greensite and P. Olesen, “Remarks on the Heavy Quark Potential in the Supergravity Approach,” hep-th/9806235. relax S. Naik, “Improved heavy quark potential at finite temperature from anti-de Sitter supergravity,” Phys. Lett. B464, 73 (1999), hep-th/9904147. relax Y. Kinar, E. Schreiber, J. Sonnenschein and N. Weiss, “Quantum fluctuations of Wilson loops from string models,” hep-th/9911123. relax A.M. Polyakov, “Quantum geometry of bosonic strings,” Phys. Lett. B103, 207 (1981); “Quantum geometry of fermionic strings,” Phys. Lett. B103, 211 (1981). relax D. Friedan, “Introduction To Polyakov’s String Theory,” in Proc. of Summer School of Theoretical Physics: Recent Advances in Field Theory and Statistical Mechanics, Les Houches, France, Aug 2-Sep 10, 1982; O. Alvarez, “Theory Of Strings With Boundaries: Fluctuations, Topology, And Quantum Geometry,” Nucl. Phys. B216, 125 (1983); H. Luckock, “Quantum Geometry Of Strings With Boundaries,” Annals Phys. 194, 113 (1989). relax A.R. Kavalov, I.K. Kostov and A.G. Sedrakian, “Dirac And Weyl Fermion Dynamics On Two-Dimensional Surface,” Phys. Lett. B175, 331 (1986); A.G. Sedrakian and R. Stora, “Dirac And Weyl Fermions Coupled To Two-Dimensional Surfaces: Determinants,” Phys. Lett. 188B, 442 (1987); D.R. Karakhanian and A.G. Sedrakian, “Heterotic string bosonization in a space of arbitrary dimension”, preprint YERPHI-1127(4)-89. relax F. Langouche and H. Leutwyler, “Two-Dimensional Fermion Determinants As Wess-Zumino Actions,” Phys. Lett. 195B, 56 (1987); “Weyl Fermions On Strings Embedded In Three-Dimensions,” Z. Phys. C36, 473 (1987); “Anomalies Generated By Extrinsic Curvature,” Z. Phys. C36, 479 (1987). relax P.B. Wiegmann, “Extrinsic Geometry Of Superstrings,” Nucl. Phys. B323, 330 (1989). relax D.R. Karakhanian, “Induced Dirac Operator And Smooth Manifold Geometry,” preprint YERPHI-1246-32-90 (1990). relax K. Lechner and M. Tonin, “The cancellation of worldsheet anomalies in the D=10 Green–Schwarz heterotic string sigma–model,” Nucl. Phys. B475, 535 (1996), hep-th/9603093. relax M.B. Green and J.H. Schwarz, Covariant description of superstrings, Phys. Lett. B136, 367 (1984); Nucl. Phys. B243, 285 (1984). relax A.M. Polyakov, unpublished (1986). relax A.M. Polyakov, “Two-dimensional Quantum gravity; Superconductivity at high $`T_c`$,” 1988 Les Houches School, “Fields, Strings and Critical Phenomena,” ed. by E. Brézin and J. Zinn-Justin, North-Holland (1990), p. 305. relax R. Kallosh and A.Y. Morozov, “Green-Schwarz Action And Loop Calculations For Superstring,” Int. J. Mod. Phys. A3, 1943 (1988). relax A.R. Kavalov and A.G. Sedrakian, “Quantum Geometry Of Covariant Superstring With N=1 Global Supersymmetry,” Phys. Lett. 182B, 33 (1986). relax L. Brink and H.B. Nielsen, A simple physical interpretation of the critical dimension of space time in dual models, Phys. Lett. B45, 332 (1973); M. Lüscher, K. Symanzik and P. Weisz, “Anomalies Of The Free Loop Wave Equation In The WKB Approximation,” Nucl. Phys. B173, 365 (1980); M. Lüscher, Symmetry breaking aspects of the roughening transition in gauge theories, Nucl. Phys. B180, 317 (1981); O. Alvarez, “The Static Potential In String Models,” Phys. Rev. D24, 440 (1981). relax P. Olesen, “Strings, Tachyons And Deconfinement,” Phys. Lett. 160B, 408 (1985). relax N. Drukker, D.J. Gross and H. Ooguri, “Wilson loops and minimal surfaces,” Phys. Rev. D60, 125006 (1999), hep-th/9904191. relax R.R. Metsaev and A.A. Tseytlin, unpublished; A.A. Tseytlin, talk at Strings’ 98, http://www.itp.ucsb.edu/online/strings98/tseytlin. relax J. Polchinski, “Evaluation Of The One Loop String Path Integral,” Commun. Math. Phys. 104, 37 (1986); G. Moore and P. Nelson, “Absence Of Nonlocal Anomalies In The Polyakov String,” Nucl. Phys. B266, 58 (1986). relax E.S. Fradkin and A.A. Tseytlin, “Quantized string models,” Ann. of Phys. 143, 413 (1982). relax A.M. Polyakov and P.B. Wiegmann, “Theory of nonabelian Goldstone bosons in two dimensions,” Phys. Lett. 131B, 121 (1983); “Goldstone Fields In Two-Dimensions With Multivalued Actions,” Phys. Lett. 141B, 223 (1984). relax W.A. Bardeen and D.Z. Freedman, “On The Energy Crisis In Anti-De Sitter Supersymmetry,” Nucl. Phys. B253, 635 (1985). relax R. Camporesi and A. Higuchi, “Stress energy tensors in anti-de Sitter space-time,” Phys. Rev. D45, 3591 (1992). relax N. Sakai and Y. Tanii, “Supersymmetry And Vacuum Energy In Anti-De Sitter Space,” Phys. Lett. 146B, 38 (1984). relax N. Sakai and Y. Tanii, “Effective Potential In Two-Dimensional Anti-De Sitter Space,” Nucl. Phys. B255, 401 (1985). relax N. Sakai and Y. Tanii, “Supersymmetry In Two-Dimensional Anti-De Sitter Space,” Nucl. Phys. B258, 661 (1985). relax T. Inami and H. Ooguri, “One Loop Effective Potential In Anti-De Sitter Space,” Prog. Theor. Phys. 73, 1051 (1985). relax C.P. Burgess and C.A. Lutken, “Propagators And Effective Potentials In Anti-De Sitter Space,” Phys. Lett. 153B, 137 (1985). relax T. Inami and H. Ooguri, “Dynamical Breakdown Of Supersymmetry In Two-Dimensional Anti-De Sitter Space,” Nucl. Phys. B273, 487 (1986). relax E.A. Ivanov and A.S. Sorin, “Wess-Zumino Model As Linear Sigma Model Of Spontaneously Broken Conformal And Osp(1,4) Supersymmetries,” Sov. J. Nucl. Phys. 30, 440 (1979); “Superfield Formulation Of Osp(1,4) Supersymmetry,” J. Phys. A13 (1980) 1159. relax J. Michelson and M. Spradlin, “Supergravity spectrum on $`AdS_2\times S^2`$,” JHEP 09, 029 (1999), hep-th/9906056. relax S.J. Avis, C.J. Isham and D. Storey, “Quantum Field Theory In Anti-De Sitter Space-Time,” Phys. Rev. D18, 3565 (1978). relax P. Breitenlohner and D.Z. Freedman, “Stability In Gauged Extended Supergravity,” Ann. Phys. 144, 249 (1982). relax C.J. Burges, D.Z. Freedman, S. Davis and G.W. Gibbons, “Supersymmetry In Anti-De Sitter Space,” Ann. Phys. 167, 285 (1986). relax C.P. Burgess, “Supersymmetry Breaking And Vacuum Energy On Anti-De Sitter Space,” Nucl. Phys. B259, 473 (1985). relax B. Allen and S. Davis, “Vacuum Energy In Gauged Extended Supergravity,” Phys. Lett. 124B, 353 (1983). relax S.M. Christensen, M.J. Duff, G.W. Gibbons and M. Rocek, “Vanishing One Loop Beta Function In Gauged $`N>4`$ Supergravity,” Phys. Rev. Lett. 45, 161 (1980). relax E. Myers, “On The Interpretation Of The Energy Of The Vacuum As The Sum Over Zero Point Energies,” Phys. Rev. Lett. 59, 165 (1987). relax D. Berenstein, R. Corrado, W. Fischler and J. Maldacena, “The operator product expansion for Wilson loops and surfaces in the large N limit,” Phys. Rev. D59, 105023 (1999), hep-th/9809188. relax R. Kallosh and J. Rahmfeld, “The GS string action on $`AdS_5\times S^5`$,” Phys. Lett. B443, 143 (1998), hep-th/9808038; I. Pesando, “A kappa gauge fixed type IIB superstring action on $`AdS_5\times S^5`$,” JHEP 9811, 002 (1998), hep-th/9808020. relax A.S. Schwarz, “The Partition Function Of A Degenerate Functional,” Commun. Math. Phys. 67, 1 (1979). relax A.S. Schwarz and A.A. Tseytlin, “Dilaton shift under duality and torsion of elliptic complex,” Nucl. Phys. B399, 691 (1993), hep-th/9210015. relax P.B. Gilkey, “The Spectral Geometry Of A Riemannian Manifold,” J. Diff. Geom. 10, 601 (1975). relax R. Camporesi, “Zeta function regularization of one loop effective potentials in anti-de Sitter space-time,” Phys. Rev. D43, 3958 (1991); R. Camporesi and A. Higuchi “Arbitrary-spin effective potentials in anti-de Sitter spacetime,” Phys. Rev. D47, 3339 (1993). relax E.S. Fradkin and A.A. Tseytlin, “Quantum String Theory Effective Action,” Nucl. Phys. B261, 1 (1985). relax I. Pesando, “The GS type IIB superstring action on $`AdS_3\times S^3\times T^4`$,” JHEP 02, 007 (1999), hep-th/9809145; J. Rahmfeld and A. Rajaraman, “The GS string action on $`AdS_3\times S^3`$ with Ramond-Ramond charge,” Phys. Rev. D60, 064014 (1999), hep-th/9809164; J. Park and S. Rey, “Green-Schwarz superstring on $`AdS_3\times S^3`$,” JHEP 01, 001 (1999), hep-th/9812062.
warning/0001/hep-ph0001022.html
ar5iv
text
# 1 Introduction ## 1 Introduction One of the main goals in $`B`$ physics is a detailed study of flavor mixing, which is encoded in the Cabibbo-Kobayashi-Maskawa (CKM) matrix of the standard model. In particular, the violation of the $`CP`$ symmetry, which the standard model describes by a nontrivial phase in the CKM matrix or equivalently by the angles of the unitarity triangle, will be investigated. Typically $`CP`$ asymmetries are expected to be large in some of the exclusive nonleptonic $`B`$ decays which, however, have only small branching ratios. Examples are the determination of $`\beta `$ from $`BJ/\psi K_s`$ and of $`\alpha `$ from $`B\pi \pi `$. In addition, in these exclusive nonleptonic decays it is very hard to obtain a good theoretical control over the hadronic uncertainties, in particular due to the presence of strong phases. On the other hand, inclusive decays have large branching fractions but typically smaller $`CP`$ asymmetries than exclusive decays . One may use parton hadron duality to obtain a good theoretical description. This has been studied by Beneke, Buchalla and Dunietz who set up a theoretically clean method to calculate the $`CP`$ asymmetries in inclusive $`B`$ decays . They still find sizable $`CP`$ asymmetries, but their measurement would require to identify charmless final states inclusively, which is not an easy task. One-particle inclusive decays lie somehow between these two cases. This class of decays still has large branching fractions and some of the expected $`CP`$ asymmetries are sizable. Furthermore, a measurement of these decays is feasible. For one-particle inclusive decays of the type B (-) D()X𝐵superscript (-) D𝑋B\to\raisebox{7.51668pt}[0.0pt]{\makebox[0.0pt][l]{\leavevmode\resizebox{9.41269pt}{2.15277pt}{\boldmath$(-)$}}}\leavevmode\hbox{$D$}^{(*)}X, a QCD based description has been developed recently, exploiting factorization and the heavy mass limit for both the $`b`$ and the $`c`$ quark . Since the expansion parameters are $`\mathrm{\Lambda }_{\mathrm{QCD}}/(m_bm_c)`$, $`1/N_C`$ and $`\alpha _s(m_c)`$, corrections to the leading term could be fairly large, in the worst case of the order of $`30\%`$. Using this method, which unfortunately is not completely model independent, we compute mixing induced time-dependent and time-integrated $`CP`$ asymmetries in the framework of the standard model. In view of the considerable uncertainties due to an unknown strong phase, our method cannot yet be used for a competitive determination of the $`CP`$ violation parameters, in particular compared to a measurement of $`\mathrm{sin}(2\beta )`$ in the “gold-plated” channel $`BJ/\psi K_s`$. However, it can be used as an estimate of the one-particle inclusive $`CP`$ asymmetries, for which we shall use present central values of the $`CP`$ angles $`\beta `$ and $`\gamma `$ . Compared to fully inclusive methods, the advantage is that we can predict asymmetries for the various spins and charges of the ground-state charmed mesons separately. This is certainly a worthwhile task, in particular since we are not aware of any previous prediction for these asymmetries, not even in the context of quark models. After introducing our notations for $`B`$ mixing in Sec. 2, we calculate the relevant matrix elements in Sec. 3 and model the form factors in Sec. 4. The numerical results are given in Sec. 5. ## 2 $`𝑪𝑷`$ asymmetries in 𝑩 (-) 𝑫()𝑿bold-→𝑩superscript (-) 𝑫𝑿B\to\raisebox{7.51668pt}[0.0pt]{\makebox[0.0pt][l]{\leavevmode\resizebox{9.41269pt}{2.15277pt}{\boldmath$(-)$}}}\leavevmode\hbox{$D$}^{(*)}X In Wigner Weisskopf approximation the time evolution of an initially pure $`B^0`$ or $`\overline{B}^0`$, $`|B_{\mathrm{phys}}^0(t)`$ $`=`$ $`g_+(t)|B^0{\displaystyle \frac{q}{p}}g_{}(t)|\overline{B}^0,`$ $`|\overline{B}_{\mathrm{phys}}^0(t)`$ $`=`$ $`g_+(t)|\overline{B}^0{\displaystyle \frac{p}{q}}g_{}(t)|B^0,`$ is determined by the time-dependent functions $`g_+(t)`$ $`=`$ $`e^{iMt\frac{1}{2}\mathrm{\Gamma }t}\left[\mathrm{cosh}{\displaystyle \frac{\mathrm{\Delta }\mathrm{\Gamma }t}{4}}\mathrm{cos}{\displaystyle \frac{\mathrm{\Delta }Mt}{2}}+i\mathrm{sinh}{\displaystyle \frac{\mathrm{\Delta }\mathrm{\Gamma }t}{4}}\mathrm{sin}{\displaystyle \frac{\mathrm{\Delta }Mt}{2}}\right]`$ $`g_{}(t)`$ $`=`$ $`e^{iMt\frac{1}{2}\mathrm{\Gamma }t}\left[\mathrm{sinh}{\displaystyle \frac{\mathrm{\Delta }\mathrm{\Gamma }t}{4}}\mathrm{cos}{\displaystyle \frac{\mathrm{\Delta }Mt}{2}}+i\mathrm{cosh}{\displaystyle \frac{\mathrm{\Delta }\mathrm{\Gamma }t}{4}}\mathrm{sin}{\displaystyle \frac{\mathrm{\Delta }Mt}{2}}\right],`$ where $`\mathrm{\Delta }M=M_HM_L>0`$ and $`\mathrm{\Delta }\mathrm{\Gamma }=\mathrm{\Gamma }_H\mathrm{\Gamma }_L<0`$ are the mass and width differences between the mass eigenstates $`|B_H>=p|B^0>+q|\overline{B}^0>`$ and $`|B_L>=p|B^0>q|\overline{B}^0>`$. The quantity $`q/p`$ is given in terms of the off-diagonal elements of the Hamiltonian $`H=Mi\mathrm{\Gamma }/2`$ of the neutral $`B`$ meson system $$\frac{q}{p}=\frac{\mathrm{\Delta }M\frac{i}{2}\mathrm{\Delta }\mathrm{\Gamma }}{2\left(M_{12}\frac{i}{2}\mathrm{\Gamma }_{12}\right)}=\frac{M_{12}^{}}{|M_{12}|}\left(1\frac{1}{2}a+𝒪(a^2)\right),a=\mathrm{Im}\left\{\frac{\mathrm{\Gamma }_{12}}{M_{12}}\right\}.$$ (3) In fact, $`\mathrm{\Gamma }_{12}/M_{12}=𝒪(m_b^2/m_t^2)`$ is very small and hence $`q/p`$ is to a good approximation a phase factor. The time-dependent rate for the decay of a $`B`$ meson into a set of final states $`|f=_i|f_i`$ can be written as $`\mathrm{\Gamma }[B(t)f]`$ $`=`$ $`{\displaystyle \frac{1}{2m_B}}{\displaystyle \underset{i}{}}{\displaystyle 𝑑\varphi _i(2\pi )^4\delta ^4(p_Bp_{f_i})B(t)\left|H_{\mathrm{eff}}\right|f_if_i\left|H_{\mathrm{eff}}\right|B(t)}`$ (4) $`=`$ $`{\displaystyle \frac{1}{2m_B}}{\displaystyle d^4xB(t)\left|H_{\mathrm{eff}}(x)\mathrm{\Pi }_fH_{\mathrm{eff}}(0)\right|B(t)},`$ where $`d\varphi _i`$ is the phase space element of the state $`|f_i`$ and $$\mathrm{\Pi }_f=\underset{i}{}𝑑\varphi _i|f_if_i|$$ (5) is the projector on the set of final states. Note that both an exclusive final state as well as inclusive states can be treated in this way. Even differential distributions can be considered if the phase spaces $`d\varphi _i`$ are not fully integrated. The $`CP`$ asymmetries we are going to consider are of the type $$𝒜_{CP}(t)=\frac{\mathrm{\Gamma }(B^0(t)f)\mathrm{\Gamma }(\overline{B}^0(t)\overline{f})}{\mathrm{\Gamma }(B^0(t)f)+\mathrm{\Gamma }(\overline{B}^0(t)\overline{f})}$$ (6) which involves the $`CP`$ conjugate set $`|\overline{f}>`$ of final states. Up to here the discussion is completely general. In the following we shall use the above formalism to compute the $`CP`$ asymmetries for one-particle inclusive final states, for which the projector reads $$\mathrm{\Pi }_f=\underset{X}{}|XYXY|,$$ (7) where $`Y`$ can be a $`D`$ or a $`\overline{D}`$ meson. Since the sum runs over all possible states $`X`$, the $`CP`$ conjugate of the projector is $$\mathrm{\Pi }_{\overline{f}}=\underset{X}{}|X\overline{Y}X\overline{Y}|.$$ (8) Inserting the time-dependent states (2) we obtain $`\mathrm{\Gamma }[B(t)YX]`$ $`=`$ $`\left|g_+(t)\right|^2\mathrm{\Gamma }_Y^{BB}+\left|{\displaystyle \frac{q}{p}}g_{}(t)\right|^2\mathrm{\Gamma }_Y^{\overline{B}\overline{B}}2\mathrm{R}\mathrm{e}\left\{{\displaystyle \frac{q}{p}}g_+^{}g_{}(t)T_Y^{B\overline{B}}\right\},`$ $`\mathrm{\Gamma }[\overline{B}(t)\overline{Y}X]`$ $`=`$ $`\left|g_+(t)\right|^2\mathrm{\Gamma }_{\overline{Y}}^{\overline{B}\overline{B}}+\left|{\displaystyle \frac{p}{q}}g_{}(t)\right|^2\mathrm{\Gamma }_{\overline{Y}}^{BB}2\mathrm{R}\mathrm{e}\left\{{\displaystyle \frac{p}{q}}g_+^{}g_{}(t)T_{\overline{Y}}^{\overline{B}B}\right\},`$ where the matrix elements are defined by $`\mathrm{\Gamma }_Y^{BB}`$ $`=`$ $`{\displaystyle \frac{1}{2m_B}}{\displaystyle d^4xB\left|H_{\mathrm{eff}}(x)\mathrm{\Pi }_YH_{\mathrm{eff}}(0)\right|B},`$ $`T_Y^{B\overline{B}}`$ $`=`$ $`{\displaystyle \frac{1}{2m_B}}{\displaystyle d^4xB\left|H_{\mathrm{eff}}(x)\mathrm{\Pi }_YH_{\mathrm{eff}}(0)\right|\overline{B}}.`$ The $`\mathrm{\Delta }B=2`$ transition matrix elements representing the interference between the mixed and the unmixed amplitudes are related by $`CPT`$ symmetry, such that $$T_Y:=T_Y^{B\overline{B}}=\left(T_Y^{\overline{B}B}\right)^{}.$$ (11) The direct $`CP`$ asymmetries in these processes are expected to be tiny. In fact, using the method described in Ref. , they turn out to be of higher order in the $`1/m`$ expansion. Hence we have $`\mathrm{\Gamma }_Y:=\mathrm{\Gamma }_Y^{BB}`$ $`=`$ $`\mathrm{\Gamma }_{\overline{Y}}^{\overline{B}\overline{B}}=\mathrm{\Gamma }(BYX),\mathrm{\Gamma }_{\overline{Y}}:=\mathrm{\Gamma }_{\overline{Y}}^{BB},`$ (12) $`T_Y`$ $`=`$ $`T_{\overline{Y}}.`$ (13) Inserting the time-dependent decay rates in Eq. (2) and neglecting both the width difference and $`a`$, such that $`q/p`$ becomes a phase factor, we obtain for the time-dependent $`CP`$ asymmetries $$𝒜_{CP}(t)=\frac{\mathrm{sin}\left(\mathrm{\Delta }Mt\right)\mathrm{Im}\left\{\frac{q}{p}T_Y\right\}}{\mathrm{cos}^2\left(\frac{\mathrm{\Delta }Mt}{2}\right)\mathrm{\Gamma }_Y+\mathrm{sin}^2\left(\frac{\mathrm{\Delta }Mt}{2}\right)\mathrm{\Gamma }_{\overline{Y}}},$$ (14) from which we get the time-integrated asymmetry $$𝒜_{CP}=\frac{2x\mathrm{Im}\left\{\frac{q}{p}T_Y\right\}}{\left(2+x^2\right)\mathrm{\Gamma }_Y+x^2\mathrm{\Gamma }_{\overline{Y}}},$$ (15) where $`x=\mathrm{\Delta }M/\mathrm{\Gamma }`$ is measured to be $`x=0.73`$ . ## 3 Transition matrix elements In order to compute the $`CP`$ asymmetries, one has to evaluate the matrix elements in Eq. (2). The total rates $`\mathrm{\Gamma }_Y`$ have already been discussed in Ref. , so we only need to calculate the interference term $`T_Y`$. The relevant pieces of the effective Hamiltonian contributing to this interference are $`(\overline{u}b)_{VA}(\overline{d}c)_{VA}`$ and $`(\overline{c}b)_{VA}(\overline{d}u)_{VA}`$ interfering with each other and $`(\overline{c}b)_{VA}(\overline{d}c)_{VA}`$ interfering with itself, so $`T_Y`$ is a sum of the two contributions $`T_Y`$ $`=`$ $`T_c+T_u,`$ (16) $`T_q`$ $`=`$ $`{\displaystyle \frac{1}{2m_B}}{\displaystyle \frac{G_F^2}{2}}V_{cb}V_{qd}^{}V_{qb}V_{cd}^{}\left|C_1\right|^2{\displaystyle \underset{X}{}}(2\pi )^4\delta ^4(p_Bp_Dp_X)`$ $`B^0\left|(\overline{q}b)_{VA}(\overline{d}c)_{VA}\right|DXDX\left|(\overline{d}q)_{VA}(\overline{c}b)_{VA}\right|\overline{B}^0.`$ Fierzing the operators into the form $`(\overline{d}b)_{VA}(\overline{u}c)_{VA}`$, $`(\overline{d}b)_{VA}(\overline{c}u)_{VA}`$ and $`(\overline{d}b)_{VA}(\overline{c}c)_{VA}`$ one can reproduce the inclusive results of Ref. . In order to evaluate the interference term for the one-particle inclusive case, we use the method developed in Ref. . It is based on factorization, which holds to leading order in the $`1/N_C`$ expansion, where $`N_C`$ is the number of QCD colors. Thus we can write the interference terms as products of two tensors $$T_q=\frac{1}{2m_B}\frac{G_F^2}{2}V_{cb}V_{qd}^{}V_{qb}V_{cd}^{}\left|C_1\right|^2\frac{d^4Q}{(2\pi )^4}K_{\mu \nu }(p_B,Q)𝑑\varphi _DP_q^{\mu \nu }(p_D,Q)$$ (18) with $`K_{\mu \nu }(p_B,Q)`$ $`=`$ $`{\displaystyle \underset{X}{}}(2\pi )^4\delta ^4(p_Bp_XQ)`$ $`B^0(p_B)\left|(\overline{d}\gamma _\mu (1\gamma _5)b)\right|XX\left|(\overline{d}\gamma _\nu (1\gamma _5)b)\right|\overline{B}^0(p_B),`$ $`P_q^{\mu \nu }(p_D,Q)`$ $`=`$ $`{\displaystyle \underset{X^{}}{}}(2\pi )^4\delta ^4(Qp_Dp_X^{})`$ $`0\left|(\overline{q}\gamma ^\mu (1\gamma _5)c)\right|D^{()}(p_D)X^{}D^{()}(p_D)X^{}\left|(\overline{c}\gamma ^\nu (1\gamma _5)q)\right|0.`$ The tensor $`K_{\mu \nu }(p_B,Q)`$ is fully inclusive and one can perform a standard short distance expansion. The resulting $`\mathrm{\Delta }B=2`$ matrix element can be parameterized by the decay constant $`f_B`$ of the $`B`$ meson and the bag factors $`B`$ and $`B_s`$ for the axial vector and the scalar current, respectively. The other tensor $`P_q^{\mu \nu }(p_D,Q)`$ involves a projection on a one-particle inclusive charmed meson state and hence we cannot perform a short distance expansion. We proceed along the same lines as in Ref. , where the rates for wrong charm decays have been modeled. Heavy quark symmetry yields the Dirac matrix structure $$P_q^{\mu \nu }(p_D,Q)\overline{H}_{D^{()}}(p_D)\gamma ^\mu (1\gamma _5)\gamma ^\nu (1\gamma _5)H_{D^{()}}(p_D),$$ (21) where the representation matrices for the charmed mesons are $$H_D=\sqrt{m_D}\frac{1+\begin{array}{c}/\hfill \\ v\hfill \end{array}_D}{2}\gamma _5,H_D^{}=\sqrt{m_D^{}}\frac{1+\begin{array}{c}/\hfill \\ v\hfill \end{array}_D^{}}{2}\begin{array}{c}/\hfill \\ ϵ\hfill \end{array}.$$ (22) In principle, all possible contractions of the light quark indices may contribute, giving rise to several form factors. For a first estimate, it is sufficient to use only the simplest one of these contractions, $$P_q^{\mu \nu }(p_D,Q)=2\pi \delta \left(\left(Qp_D\right)^2m_q^2\right)\mathrm{Tr}\left\{\begin{array}{c}/\hfill \\ p\hfill \end{array}_D\gamma ^\mu (1\gamma _5)\left(\begin{array}{c}/\hfill \\ Q\hfill \end{array}\begin{array}{c}/\hfill \\ p\hfill \end{array}_D\right)\gamma ^\nu (1\gamma _5)\right\}\stackrel{~}{f}_{qY},$$ (23) corresponding to a replacement of the $`D^{()}X`$ final state by a pair of free quarks, rescaled by an operator- and decay-channel-specific form factor $`\stackrel{~}{f}_{qY}`$, where $`Y`$ is one of the ground state $`D`$ mesons. In the following, we call this contraction “partonic.” Using this ansatz and the heavy mass limit, the transition matrix elements read $`T_c`$ $`=`$ $`{\displaystyle \frac{G_F^2m_B^3f_B^2}{24\pi }}(V_{cb}V_{cd}^{})^2\left|C_1\right|^2\sqrt{14z}\left[(14z)B+2(1+2z)B_S\right]\stackrel{~}{f}_{cY},`$ (24) $`T_u`$ $`=`$ $`{\displaystyle \frac{G_F^2m_B^3f_B^2}{24\pi }}V_{cb}V_{ud}^{}V_{ub}V_{cd}^{}\left|C_1\right|^2(1z)^2\left[(1z)B+2(1+2z)B_S\right]\stackrel{~}{f}_{uY},`$ (25) where $`z=(m_c/m_b)^2`$ and $`C_1`$ is the Wilson coefficient of the effective Hamiltonian in the notation of Ref. . Equations (24) and (25) correspond to the expression for the width difference of neutral heavy meson systems . In the standard CKM parametrization, the phases of the transition matrix elements are $`\mathrm{arg}(T_c)`$ $`=`$ $`0,`$ (26) $`\mathrm{arg}(T_u)`$ $`=`$ $`\mathrm{arg}(V_{ub})=\gamma ,`$ (27) $`\mathrm{arg}(q/p)`$ $`=`$ $`\mathrm{arg}(V_{td}^2)=2\beta ,`$ (28) such that $$\mathrm{Im}\left\{\frac{q}{p}T_Y\right\}=\mathrm{sin}(2\beta )\left|T_c\right|+\mathrm{sin}(2\beta +\gamma )\left|T_u\right|.$$ (29) ## 4 Modeling the form factors We assume that the form factors $`\stackrel{~}{f}_{qY}`$ do not vary strongly over the accessible phase space and hence we approximate them by constants. For the case $`q=c`$, these constants have been fitted to the wrong charm yield in $`B`$ decays . Operators analogous to the case $`q=u`$ are Cabibbo suppressed when calculating wrong charm rates, so they did not appear in Ref. . Assuming that all charm quarks eventually hadronize to $`D`$ mesons, we use $$\stackrel{~}{f}_{uD^0}+\stackrel{~}{f}_{uD^+}=1.$$ (30) To resolve the spin and charge counting, we first discuss the heavy mass limit where the pseudoscalar and vector charmed mesons form a degenerate ground state doublet. The decay of vector to pseudoscalar mesons will be discussed below. In the following, $`D_{\mathrm{dir}}`$ refers to those $`D`$ mesons that do not result from $`D^{}`$ decays, and $`D^{()}`$ can be either $`D_{\mathrm{dir}}`$ or $`D^{}`$. As long as the light quark spin indices of the $`D^{()}`$ meson representation matrices are contracted with each other, Eq. (21) reproduces the naive spin counting $$\stackrel{~}{f}_{qD^0}=3\stackrel{~}{f}_{qD_{\mathrm{dir}}^0},\stackrel{~}{f}_{qD^+}=3\stackrel{~}{f}_{qD_{\mathrm{dir}}^+}.$$ (31) Different contractions yield results of comparable size. The experimental spin counting factor appears to be smaller by roughly a factor of two . Since this effect is not yet understood, we treat it as an uncertainty. Concerning charge counting, we argued by isospin symmetry that in the case $`q=c`$ we have $$\stackrel{~}{f}_{cD^{()0}}=\stackrel{~}{f}_{cD^{()+}}.$$ (32) In the case $`q=u`$, two topologies can contribute to the decay amplitude: the charm quark can either hadronize with the $`u`$ quark from the weak effective current, in which case the isospin of the state $`|X`$ is $`I_X=0`$, or with a $`u`$ or $`d`$ quark from vacuum, which contains both $`I_X=0`$ and $`I_X=1`$ contributions. In the case $`I_X=0`$, both amplitudes can interfere, so there are three contributions to the decay rate $`\stackrel{~}{f}_{uD^{()0}}`$ $`=`$ $`\left|a_1+a_2\right|^2=\left|a_1\right|^2+\left|a_2\right|^2+2\mathrm{R}\mathrm{e}\left\{a_1^{}a_2\right\}`$ $`\stackrel{~}{f}_{uD^{()+}}`$ $`=`$ $`\left|a_2\right|^2,`$ see Figs. 33. One might doubt whether using the partonic contraction given in Eq. (23) is justified for all the topologies, as it appears to correspond to the topology in Fig. 3, while the topology in Fig. 3 should rather be described by the contraction $$P_q^{\mu \nu }(p_D,Q)\mathrm{Tr}\left\{\overline{H}_{D^{()}}(p_D)\gamma ^\mu (1\gamma _5)\right\}\mathrm{Tr}\left\{\gamma ^\nu (1\gamma _5)H_{D^{()}}(p_D)\right\}.$$ (34) This is not a problem for three reasons. First, we do not claim to be able to accurately model the matrix element, but we only give the simplest possible ansatz by rescaling the partonic result. In particular, it is clearly not yet feasible to model particular contributions individually. We only use the three topologies to estimate the integrated relative magnitudes of the two main contributions and to bound the magnitude of their interference term. Secondly, neither the time-dependent nor the time-integrated asymmetries depend on the choice of the contraction unless studied differentially in the momentum of the charmed meson, which so far we do not attempt to do. Finally, as noted in Ref. , the choice of the wrong charm contraction appeared to have little influence even on differential observables. The topologies in Figs. 3 and 3 also occur in wrong charm production in $`B`$ decays. Figure 3 corresponds to the process $`BD_s^{()+}X`$, Fig. 3 to the process $`BD^{()}X`$, where $`D^{()}`$ can be either $`D^{()0}`$ or $`D^{()+}`$. Both contributions are experimentally known to be of similar size, i.e., $`(10\pm 2.5)\%`$ and $`(7.9\pm 2.2)\%`$ , respectively, such that $$\left|a_1\right|^2=2\left|a_2\right|^2.$$ (35) The relative phase of the two contributions is unknown. Therefore, although it may be large, we have to treat the interference part as a theoretical uncertainty. This is acceptable since the $`q=u`$ contribution is smaller than the $`q=c`$ contribution according to $$\left|\frac{T_u}{T_c}\right|=\left|\frac{V_{ub}}{V_{cb}}\frac{V_{ud}}{V_{cd}}\right|\frac{(1z)^2(1+z)}{\sqrt{14z}}\frac{\stackrel{~}{f}_u}{\stackrel{~}{f}_c}\left|\frac{V_{ub}}{V_{cb}}\right|\frac{(1+z)}{\left|V_{cd}\right|}0.4.$$ (36) Off the heavy mass limit, $`D^{}D`$ decay has to be taken into account. In the same way as in Ref. , we get $`\stackrel{~}{f}_{qD^+}`$ $`=`$ $`\stackrel{~}{f}_{qD_{\mathrm{dir}}^+}+\mathrm{Br}\left(D^+D^+X\right)\stackrel{~}{f}_{qD^+}`$ $`\stackrel{~}{f}_{qD^0}`$ $`=`$ $`\stackrel{~}{f}_{qD_{\mathrm{dir}}^0}+\stackrel{~}{f}_{qD^0}+\mathrm{Br}\left(D^+D^0X\right)\stackrel{~}{f}_{qD^+}.`$ The coefficients obtained from Eqs. (30)–(4) and Ref. are summarized in Table 1. The ranges given result from varying the spin counting factor in Eq. (31) from $`3`$ down to $`3/2`$ and the interference in Eq. (4) from the central value of vanishing interference to full constructive and destructive interference. ## 5 Results We have computed the parameters for the time-dependent $`CP`$ asymmetries as well as the time-integrated asymmetries. We have inserted recent values for $`\mathrm{sin}2\beta =0.75`$ and $`\gamma =68^{}`$ . In addition, we use $`V_{cb}=0.04`$, $`V_{ub}=0.08V_{cb}`$, $`z=0.09`$, $`x=0.73`$, $`f_B=180`$ MeV, $`\mathrm{Br}\left(D^+D^0Y\right)=1\mathrm{Br}\left(D^+D^+Y\right)=0.683`$ and $`C_1=B=B_S=1`$. The results of the calculations can be found in Fig. 4 and Table 2. To assess the uncertainties involved in Fig. 4, note that according to Eq. (14) the shapes of the time-dependent asymmetries are determined by the ratios of the wrong to right charm rates $`\mathrm{\Gamma }_{\overline{Y}}/\mathrm{\Gamma }_Y`$. We checked numerically that the shapes would hardly change even if these ratios were off by $`30\%`$. The dominant contribution to the uncertainty of the amplitudes arises from the transition matrix elements $`T_Y`$ and is directly proportional to the uncertainties of the time-integrated asymmetries given in Table 2. Suppose $`N`$ perfectly tagged $`B^0`$ decays are recorded in an experiment. In order to establish the asymmetry in a channel with a branching ratio $`b`$ on the $`3\sigma `$ level, $$\frac{𝒜}{3}\mathrm{\Delta }𝒜=\frac{1}{\sqrt{2bN}}$$ (38) has to be satisfied. The necessary numbers of tagged $`B^0`$ decays are given in the last column of Table 2. Since the asymmetry tends to be roughly inversely proportional to the branching ratio by Eq. (15), we obtain from Eq. (38) $$N\frac{1}{𝒜^2b}b,$$ (39) such that rare channels are advantageous for observing one-particle inclusive asymmetries. The channel $`B^0\overline{D}^0X`$ deserves a further comment. Looking at Fig. 4, there is an obvious problem at small proper decay times. The reason for this problem is that we have discussed all the rates only to leading order in the combined $`1/N_C`$ and $`1/m_Q`$ expansions. However, this leading term vanishes for the channel $`B^0\overline{D}^0X`$ and thus subleading terms become relevant. On the other hand, the numerator $`T_Y`$ of the $`CP`$ asymmetries is given by a matrix element of a dimension six operator and hence is suppressed compared to the leading terms of most of the rates. In other words, while in most of the rates the asymmetries are of subleading order $`f_B^2/m_B^2`$, this is not the case for the channel $`B^0\overline{D}^0X`$. Unfortunately we cannot compute this possibly large asymmetry, since this would involve to compute subleading terms for the decay rate. Hence we try to estimate the asymmetry by varying $`\mathrm{Br}\left(B^0\overline{D}^0X\right)`$ in Eq. (15) and show the reaction of the asymmetry in Fig. 6 and of the necessary number of tagged $`B^0`$ events in Fig. 6. The wrong charm asymmetry is practically unaffected by $`\mathrm{Br}\left(B^0\overline{D}^0X\right)`$ since the pole occurs near four average lifetimes where most of the $`B`$ mesons have already decayed, but the right charm asymmetry turns out to be extremely sensitive. Therefore we cannot predict the latter quantitatively, but it can be as large as several percent, and it will be measurable with a few $`\mathrm{100\hspace{0.17em}000}`$ tagged $`B^0`$ events. ## 6 Conclusion Motivated by the work on fully inclusive $`CP`$ asymmetries and the question how to measure them, we studied one-particle inclusive $`CP`$ asymmetries. In the final state only a (-) D()superscript (-) D\raisebox{7.51668pt}[0.0pt]{\makebox[0.0pt][l]{\leavevmode\resizebox{9.41269pt}{2.15277pt}{\boldmath$(-)$}}}\leavevmode\hbox{$D$}^{(*)} meson has to be identified and thus they are experimentally more easily accessible than the fully inclusive $`CP`$ asymmetries. We have used a similar method as in in Ref. to calculate the time-dependent and time-integrated $`CP`$ asymmetries for one-particle inclusive B (-) D()X𝐵superscript (-) D𝑋B\to\raisebox{7.51668pt}[0.0pt]{\makebox[0.0pt][l]{\leavevmode\resizebox{9.41269pt}{2.15277pt}{\boldmath$(-)$}}}\leavevmode\hbox{$D$}^{(*)}X decays. It turns out that, as in Ref. , one cannot avoid to introduce some model dependence. Furthermore, there is also some dependence on an unknown relative phase, which we treat as an uncertainty. Due to these uncertainties we cannot expect our method to compete with proposed methods using “gold-plated” channels for determining CKM parameters, but we can still give estimates for the expected $`CP`$ asymmetries of the different ground state $`\mathbf{(}\mathbf{}\mathbf{)}`$ $`D`$ mesons. For most of the asymmetries we find results of a few $`10^3`$, but some are expected to be as large as several percent. These effects should be observable at the $`B`$ factories. The channels involving right and wrong charm neutral vector mesons turn out to be most promising: they are expected to have the largest asymmetries, and the theoretical method yields the best results for the production rates and spectra of the vector mesons . ## Acknowledgments The authors thank Thomas Gehrmann for fruitful discussions. This work (X. C. during his time in Karlsruhe, T. M. and I. S.) was supported by the DFG Graduiertenkolleg “Elementarteilchenphysik an Beschleunigern” and by the DFG Forschergruppe “Quantenfeldtheorie, Computeralgebra und Monte-Carlo-Simulation.”
warning/0001/hep-ex0001032.html
ar5iv
text
# Limit on a horizontal emittance in high energy muon colliders due to synchrotron radiation. Talk at the Workshop Studies on Colliders and Collider Physics at the Highest Energies: Muon Colliders at 10 TeV to 100 TeV, 27 September - 1 October, 1999 Montauk, New York, USA, be published by the American Institute of Physics. ## Abstract It is shown that at a 100 TeV muon collider the synchrotron radiation in the ring will determine the minimum horizontal emittance. In all e<sup>+</sup>e<sup>-</sup> storage ring the horizontal emittance is determined by quantum nature of the synchrotron radiation. Emission of photons in regions with non-zero dispersion leads to growth of the horizontal emittance while the average energy loss leads to the cooling. As the result, some equilibrium emittance is reached. This is well known and corresponding formulae can be found elsewhere . At muon colliders, the synchrotron radiation power is much smaller than that at e<sup>+</sup>e<sup>-</sup> storage rings due to higher particle mass: $`P(EB)^2/m^4`$; however this suppression is compensated by the much higher energy and magnetic field. In general, emittance in a storage ring depends on time in the following way $$ϵ_x=ϵ_D(1e^{t/\tau })+ϵ_0e^{t/\tau },$$ (1) where $`ϵ_0`$ is the initial emittance, $`ϵ_D`$ is the equilibrium emittance, $`\tau `$ is the damping time. One can check (see also the B.King’s table) that for the 2E=100 TeV muon collider $`\gamma _\mu \tau _\mu \tau _D`$ 1 sec, so the emittance will be close to the equilibrium one. The equilibrium normalized ($`ϵ_n=\gamma ϵ`$) horizontal emittance in a damping ring $$ϵ_D=\frac{55\mathrm{}}{32\sqrt{3}mc}\frac{\gamma ^3}{J_x}\frac{H}{\rho },$$ (2) where $`J_x1`$ is the damping partition number, $`H\beta ^3/\rho ^2`$ is the average value of the “H-function” characterizing a storage ring magnetic structure. One can see that for fixed energy and magnetic field $`ϵ_{nx}1/m.^4`$ For an optimized FODO ring structure of the muon collider the minimum normalized horizontal emittance due to synchrotron radiation $$ϵ_D100\frac{E^3[\text{GeV}]}{N^3}\left(\frac{m_e}{m_\mu }\right)^4\text{cm rad},$$ (3) where $`N`$ is the number of bending magnets in the ring. Extrapolating the number of magnets in the current designs to the 100 TeV energy as $`N\sqrt{E}`$ (the usual energy dependence of the $`\beta `$-function) one can get $`N1000`$ for $`2E=100`$ TeV. For this structure we obtain the normalized horizontal emittance at $`t=\tau _D`$ $$ϵ_{nx}0.63ϵ_D5\times 10^3\text{cm rad}.$$ (4) For comparison, in the B.King’s table $`ϵ_{nx}=8.7\times 10^4`$ cm rad for “evolutionary” extrapolation and $`ϵ_{nx}=2.1\times 10^5`$ cm rad for “ultra-cold beams”, which is smaller than above result by factors of 5 and 250, respectively. This example shows that the considered effect is important for high energy muon colliders and this problem needs more accurate consideration. One should consider the more realistic case of the isochronous ring (which is necessary for muon colliders), including W-shielding which reduces the quad strength (and correspondingly $`N`$).
warning/0001/nucl-th0001054.html
ar5iv
text
# Cranked Relativistic Hartree-Bogoliubov Theory: Formalism and Application to the Superdeformed Bands in the 𝐴∼190 region ## I Introduction The development of self-consistent microscopic many-body mean field theories aimed on the description of low-energy nuclear phenomena provides necessary theoretical tools for an exploration of the nuclear chart into know and unknown regions. This development is motivated by theoretical and experimental reasons. Compared with conventional approaches such as, for example, the macroscopic+microscopic method, self-consistent theories rely on a smaller number of assumptions and start from a more microscopic level. For example, the starting point of non-relativistic mean field theories is an effective interaction between nucleons constituting the nucleus. Such theories based either on zero range Skyrme forces or finite range Gogny forces have been widely used starting from seventies. The next step is replacing the Schrödinger equation by the Dirac equation and thus considering the relativistic mean field (RMF) theory . In RMF theory, the nucleus is described as a system of point-like nucleons, Dirac spinors, which interact in a phenomenological way by the exchange of mesons, such as the $`\sigma `$-meson responsible for the large scalar attraction at intermediate distances, the $`\omega `$-meson for the vector repulsion at short distances and the $`\rho `$-meson for the asymmetry properties of nuclei with large neutron or proton excess. Such a description has a clear advantage that the spin-orbit splitting, which plays an extremely important role in low-energy nuclear physics, emerges in a natural way as a genuine relativistic effect. In addition, the pseudo-spin symmetry, the origin of which was a long-standing puzzle, does find a natural explanation in the framework of RMF theory . RMF theory has been extremely successful also in the description of many other facets of low-energy nuclear physics, such as ground state properties, giant resonances, superdeformed rotating nuclei etc., see Ref. for an overview. Since the discovery of the first superdeformed rotational (SD) band in <sup>152</sup>Dy , the investigation of superdeformation at high angular momentum remains one of the most challenging topics of nuclear structure. The rich variety of physical phenomena at superdeformed shapes is based on a complicated and rather subtle interplay of collective and single-particle properties. Although at the present stage, a general understanding of this phenomenon has been achieved, there are still many unresolved questions related, for example, to the phenomenon of identical bands and to the treatment of pairing correlations. In addition, the microscopic theoretical models used so far are still far from a precise quantitative description of SD bands which indicates the necessity of further improvements. Different theoretical methods based mainly on the concept of the cranking model of Inglis have been employed for the quantitative description of various high-spin phenomena at SD shapes, see for example recent overviews in Refs. . The cranked version of the RMF theory - the Cranked Relativistic Mean Field (CRMF) theory is amongst the most successful ones. It has been applied in a systematic way to the description of SD bands in different mass regions such as $`A60`$ , $`A80`$ and $`A140150`$ . Pairing correlations are expected to be considerably quenched in SD bands of these regions at high spin and thus they have been neglected in all studies quoted above. One should clearly recognize that the neglect of pairing correlations is an approximation because pairing correlations being weak are still present even at the highest rotational frequencies. Despite this a very successful description of many properties of SD bands, such as dynamic $`J^{(2)}`$ and kinematic $`J^{(1)}`$ moments of inertia, absolute and relative charge quadrupole moments, effective alignments $`i_{eff}`$, single-particle properties in the SD minimum etc., has been obtained in these studies in an unpaired formalism. However, the rotational properties of nuclei at low and medium spin are strongly affected by pairing correlations. In order to describe such properties within the relativistic framework, we have developed the Cranked Relativistic Hartree-Bogoliubov (CRHB) theory. This theory is an extension of CRMF theory to the description of pairing correlations in rotating nuclei. The brief outline of this theory and its application to the study of several yrast SD bands observed in the $`A190`$ mass region has been reported in Ref. . The present manuscript represents an extension of this investigation where both the theoretical formalism and the calculations will be presented in much greater details. The paper is organized in the following way: In Section II a detailed description of the CRHB theory without and with approximate particle number projection by means of the Lipkin-Nogami method and some of the specific features of the present calculations are presented. The shell structure in the $`A190`$ mass region of superdeformation and the impact of pairing and particle number projection on the rotational and deformation properties of rotating nuclei are investigated in detail on the example of the lowest SD bands in <sup>192</sup>Hg and <sup>194</sup>Pb nuclei in Section III. In Section IV, we study the dependence of the results of CRHB calculations with particle number projection on the parametrization of the RMF Lagrangian and the Gogny force using SD bands in <sup>194</sup>Hg and <sup>194</sup>Pb as an example. The properties of yrast SD bands observed so far in even-even nuclei of the $`A190`$ mass region are systematically studied in Section V. Finally, Section VI summarizes our main conclusions. ## II Cranked Relativistic Hartree-Bogoliubov (CRHB) Theory ### A The CRHB equations In relativistic mean field (RMF) theory the nucleus is described as a system of point-like nucleons, Dirac spinors, coupled to mesons and to the photons. The nucleons interact by the exchange of several mesons, namely a scalar meson $`\sigma `$ and three vector particles $`\omega `$, $`\rho `$ and the photon. The isoscalar-scalar $`\sigma `$-mesons provide a strong intermediate range attraction between the nucleons. For the three vector particles we have to distinguish the time-like components and the spatial components. For the photons this means the Coulomb field and possible magnetic field in the case where currents play a role. For the isoscalar-vector $`\omega `$-meson the time-like component provides a very strong repulsion at short distances for all combinations of particles, $`pp`$, $`nn`$ and $`pn`$. For the isovector-vector $`\rho `$-meson the time-like components give rise to a short range repulsion for like particles ($`pp`$ and $`nn`$) and a short range attraction for unlike particles ($`np`$). They also have a strong influence on the symmetry energy. In addition, the spatial components of the $`\omega `$ and $`\rho `$-mesons lead to an interaction between possible currents, which are for the $`\omega `$-meson attractive for all combinations ($`pp`$, $`nn`$ and $`pn`$-currents) and for the $`\rho `$-meson attractive for $`pp`$ and $`nn`$-currents but repulsive for $`pn`$-currents. We have to keep in mind, however, that within mean field theory these currents only occur in cases of time-reversal breaking mean fields as, for instance, in the case of the Coriolis fields. The starting point of RMF theory is the well known local Lagrangian density $``$ $`=`$ $`\overline{\psi }(i\gamma ^\mu _\mu m)\psi +{\displaystyle \frac{1}{2}}(_\mu \sigma ^\mu \sigma m_\sigma ^2\sigma ^2){\displaystyle \frac{1}{3}}g_2\sigma ^3{\displaystyle \frac{1}{4}}g_3\sigma ^4`$ (3) $`{\displaystyle \frac{1}{4}}\mathrm{\Omega }_{\mu \nu }\mathrm{\Omega }^{\mu \nu }+{\displaystyle \frac{1}{2}}m_\omega ^2\omega _\mu \omega ^\mu {\displaystyle \frac{1}{4}}\stackrel{}{R}_{\mu \nu }\stackrel{}{R}^{\mu \nu }+{\displaystyle \frac{1}{2}}m_\rho ^2\stackrel{}{\rho }_\mu \stackrel{}{\rho }^\mu {\displaystyle \frac{1}{4}}F_{\mu \nu }F^{\mu \nu }`$ $`g_\sigma \overline{\psi }\sigma \psi g_\omega \overline{\psi }\gamma ^\mu \omega _\mu \psi g_\rho \overline{\psi }\gamma ^\mu \stackrel{}{\tau }\stackrel{}{\rho }_\mu \psi e\overline{\psi }\gamma ^\mu {\displaystyle \frac{1\tau _3}{2}}A_\mu \psi ,`$ where the non-linear self-coupling of the $`\sigma `$-field, which is important for an adequate description of nuclear surface properties and the deformations of finite nuclei, is taken into account according to Ref. . The Lagrangian (3) contains as parameters the masses of the mesons $`m_\sigma `$, $`m_\omega `$ and $`m_\rho `$, the coupling constants $`g_\sigma `$, $`g_\omega `$ and $`g_\rho `$ and the non-linear terms $`g_2`$ and $`g_3`$. The field tensors for the vector mesons and the photon field are: $`\mathrm{\Omega }_{\mu \nu }=_\mu \omega _\nu _\nu \omega _\mu ,\stackrel{}{R}_{\mu \nu }=_\mu \stackrel{}{\rho }_\nu _\nu \stackrel{}{\rho }_\mu ,F_{\mu \nu }=_\mu A_\nu _\nu A_\mu .`$ (4) In the present state of the art of the RMF theory, the meson and photon fields are treated as classical fields. Two approximations, namely the mean-field approximation, in which the meson field operators are replaced by their expectation values $`\sigma =\sigma _0,\omega _\mu =\delta _{\mu 0}\omega _0`$ (5) which is justified by the fact that the source terms are large , and the No-sea approximation, in which only positive-energy states are taken into account , are employed in order to solve the Lagrangian (3). For classical fields we can restrict ourselves to the intrinsic symmetry violating product wave function $`|\mathrm{\Phi }`$ which can be represented as a generalized Slater determinant. As long as we consider time-independent static or quasistatic fields, this limits the applicability of CRHB theory to the yrast bands and to the excited bands of dominantly quasiparticle nature. In particular, the bands based on low-lying collective vibrations, such as for example $`\gamma `$-, $`\beta `$-vibrational bands, cannot be described in the present formalism since they have (for example, in even-even nuclei) an intrinsic wave function which is a linear superposition of many two-quasiparticle states. In order to study such bands one should employ time-dependent fields in the random phase approximation in the rotating frame: the task which has so far been resolved only in simple non-relativistic models with separable forces . As discussed in detail in Refs. in the unpaired formalism, the transformation to the rotating frame within the framework of the cranking model leads to the cranked relativistic mean field equations. Note that in the present investigation we restrict ourselves to one-dimensional rotation with rotational frequency $`\mathrm{\Omega }_x`$ around the $`x`$-axis. Since pairing correlations work only between the fermions, the Bogoliubov transformation affects directly only the fermionic part. Thus the time-independent inhomogeneous Klein-Gordon equations for the mesonic fields obtained by means of variational principle are given in the CRHB theory by $`\left\{\mathrm{\Delta }(\mathrm{\Omega }_x\widehat{L}_x)^2+m_\sigma ^2\right\}\sigma (𝒓)`$ $`=`$ $`g_\sigma \rho _s(𝒓)`$ (6) $`g_2\sigma ^2(𝒓)g_3\sigma ^3(𝒓),`$ (7) $`\left\{\mathrm{\Delta }(\mathrm{\Omega }_x\widehat{L}_x)^2+m_\omega ^2\right\}\omega _0(𝒓)`$ $`=`$ $`g_\omega \rho _v^{is}(𝒓),`$ (8) $`\left\{\mathrm{\Delta }[\mathrm{\Omega }_x(\widehat{L}_x+\widehat{S}_x)]^2+m_\omega ^2\right\}𝝎(𝒓)`$ $`=`$ $`g_\omega 𝒋^{is}(𝒓),`$ (9) $`\left\{\mathrm{\Delta }(\mathrm{\Omega }_x\widehat{L}_x)^2+m_\rho ^2\right\}\rho _0(𝒓)`$ $`=`$ $`g_\rho \rho _v^{iv}(𝒓),`$ (10) $`\left\{\mathrm{\Delta }[\mathrm{\Omega }_x(\widehat{L}_x+\widehat{S}_x)]^2+m_\rho ^2\right\}𝝆(𝒓)`$ $`=`$ $`g_\rho 𝒋^{iv}(𝒓),`$ (11) $`\mathrm{\Delta }A_0(𝒓)`$ $`=`$ $`e\rho _v^p(𝒓),`$ (12) $`\mathrm{\Delta }𝑨(𝒓)`$ $`=`$ $`e𝒋^p(𝒓),`$ (13) where the source terms are sums of bilinear products of baryon amplitudes $`\rho _s(𝒓)`$ $`=`$ $`{\displaystyle \underset{k>0}{}}[V_k^n(𝒓)]^{}\widehat{\beta }V_k^n(𝒓)+[V_k^p(𝒓)]^{}\widehat{\beta }V_k^p(𝒓),`$ (14) $`\rho _v^{is}(𝒓)`$ $`=`$ $`{\displaystyle \underset{k>0}{}}[V_k^n(𝒓)]^{}V_k^n(𝒓)+[V_k^p(𝒓)]^{}V_k^p(𝒓),`$ (15) $`\rho _v^{iv}(𝒓)`$ $`=`$ $`{\displaystyle \underset{k>0}{}}[V_k^n(𝒓)]^{}V_k^n(𝒓)[V_k^p(𝒓)]^{}V_k^p(𝒓),`$ (16) $`𝒋^{is}(𝒓)`$ $`=`$ $`{\displaystyle \underset{k>0}{}}[V_k^n(𝒓)]^{}\widehat{𝜶}V_k^n(𝒓)+[V_k^p(𝒓)]^{}\widehat{𝜶}V_k^p(𝒓),`$ (17) $`𝒋^{iv}(𝒓)`$ $`=`$ $`{\displaystyle \underset{k>0}{}}[V_k^n(𝒓)]^{}\widehat{𝜶}V_k^n(𝒓)[V_k^p(𝒓)]^{}\widehat{𝜶}V_k^p(𝒓).`$ (18) The sums over $`k>0`$ run over all quasiparticle states corresponding to positive energy single-particle states (no-sea approximation). In Eqs. (13,18), the indexes $`n`$ and $`p`$ indicate neutron and proton states, respectively, and the indexes $`is`$ and $`iv`$ are used for isoscalar and isovector quantities. $`\rho _v^p(𝒓)`$, $`𝒋^p(𝒓)`$ in Eq. (13) correspond to $`\rho _v^{is}(𝒓)`$ and $`𝒋^{is}(𝒓)`$ defined in Eq. (18), respectively, but with the sums over neutron states neglected. Note that the Coriolis term for the Coulomb potential $`A_0(𝒓)`$ and the spatial components of the vector potential $`𝑨(𝒓)`$ are neglected in Eqs. (13) since the coupling constant of the electromagnetic interaction is small compared with the coupling constants of the meson fields. On this classical level only the direct terms of the potentials are taken into account since most parametrizations of the RMF Lagrangian have been fitted to experimental data neglecting exchange terms. This means that the exchange terms are not fully neglected, but taken into account in an averaged way by adjusting the parameters of the direct terms. The comparison of Eqs. (13,18) with Eqs. (4) and (5) in Ref. indicates that the pairing correlations between the fermions have also an impact on mesonic fields through the redefinition of the various nucleonic densities and currents. While the occupation probabilities of the single-nucleon orbitals are equal 1 and 0 for the orbitals below and above the Fermi level in the systems with no pairing, they are between 0 and 1 when the pairing between the fermions is taken into account. Contrary to the applications of the Hartree-(Fock)-Bogoliubov theory for the ground states of even-even nuclei, the time-reversal symmetry of the intrinsic wave function $`|\mathrm{\Phi }`$ is broken by the Coriolis operator $`\mathrm{\Omega }_x\widehat{J}_x`$ ($`\widehat{J_x}`$ is the projection of total angular momentum on the rotation axis) in the case of rotating nuclei. Therefore, we do not know a priori the conjugate states in the canonical basis and thus one must solve the full Hartree-Bogoliubov problem in the rotating frame . The solution of this problem in the RMF theory is to large extent similar to the one obtained earlier in the non-relativistic case as far as we treat pairing in non-relativistic fashion as it is done in the CRHB theory. Thus in the following discussion we will mainly concentrate on the features specific for relativistic case. The CRHB equations for the fermions in the rotating frame are given in one-dimensional cranking approximation by $`\left(\begin{array}{cc}\widehat{h}_D\lambda _\tau \mathrm{\Omega }_x\widehat{J}_x& \widehat{\mathrm{\Delta }}\\ \widehat{\mathrm{\Delta }}^{}& \widehat{h}_D^{}+\lambda _\tau +\mathrm{\Omega }_x\widehat{J}_x^{}\end{array}\right)\left(\begin{array}{c}U_k(𝒓)\\ V_k(𝒓)\end{array}\right)=E_k\left(\begin{array}{c}U_k(𝒓)\\ V_k(𝒓)\end{array}\right)`$ (19) where $`\widehat{h}_D`$ is the Dirac Hamiltonian for the nucleon with mass $`m`$ $`\widehat{h}_D=𝜶(i\mathbf{}𝑽(𝒓))+V_0(𝒓)+\beta (m+S(𝒓))`$ (20) and $`\lambda _\tau `$ are the chemical potentials defined from the average particle number constraints for protons and neutrons ($`\tau =p,n`$) $`\mathrm{\Phi }_{\mathrm{\Omega }_x}|\widehat{N}_p|\mathrm{\Phi }_{\mathrm{\Omega }_x}=Z,\mathrm{\Phi }_{\mathrm{\Omega }_x}|\widehat{N}_n|\mathrm{\Phi }_{\mathrm{\Omega }_x}=N.`$ (21) The particle number expectation values $`\mathrm{\Phi }_{\mathrm{\Omega }_x}|\widehat{N}_\tau |\mathrm{\Phi }_{\mathrm{\Omega }_x}`$ are defined via the (normal) density matrices $`\rho _p`$ and $`\rho _n`$ $`\mathrm{\Phi }_{\mathrm{\Omega }_x}|\widehat{N}_\tau |\mathrm{\Phi }_{\mathrm{\Omega }_x}=Tr(\rho _\tau )\mathrm{where}\rho _\tau =V_\tau ^{}V_\tau ^T.`$ (22) The Dirac Hamiltonian contains an attractive scalar potential $`S(𝒓)`$ $`S(𝒓)=g_\sigma \sigma (𝒓),`$ (23) a repulsive vector potential $`V_0(𝒓)`$ $`V_0(𝒓)=g_\omega \omega _0(𝒓)+g_\rho \tau _3\rho _0(𝒓)+e{\displaystyle \frac{1\tau _3}{2}}A_0(𝒓),`$ (24) a magnetic potential $`𝑽(𝒓)`$ $`𝑽(𝒓)=g_\omega 𝝎(𝒓)+g_\rho \tau _3𝝆(𝒓)+e{\displaystyle \frac{1\tau _3}{2}}𝑨(𝒓),`$ (25) and a Coriolis term $`\mathrm{\Omega }_x\widehat{J}_x=\mathrm{\Omega }_x(\widehat{L}_x+{\displaystyle \frac{1}{2}}\widehat{\mathrm{\Sigma }}_x).`$ (26) The latter two terms are the contributions to the mean field which break time-reversal symmetry and induce currents. In the Dirac equation the field $`𝑽(𝒓)`$ has the structure of a magnetic field. Time-reversal symmetry is broken when the orbitals which are time-reversal counterparts are not occupied pairwise. At no rotation, this usually takes place in one-(multi-)quasiparticle configurations where the magnetic potential $`𝑽`$($`𝒓`$) breaks time-reversal symmetry. In rotating nuclei the time-reversal symmetry is additionally broken by the Coriolis field. The nuclear currents of Eq. (13) are the sources for the space-like components of the vector $`𝝎(𝒓)`$, $`𝝆(𝒓)`$ and $`𝑨(𝒓)`$ fields which give rise to polarization effects in the Dirac spinors through the magnetic potential $`𝑽(𝒓)`$ of Eq. (25). This effect is commonly referred to as nuclear magnetism . It turns out that it is very important to take it into account for the proper description of currents, magnetic moments and moments of inertia . In the present calculations the spatial components of the vector mesons are properly taken into account in a fully self-consistent way. In Eq. (19), the rotational frequency $`\mathrm{\Omega }_x`$ along the $`x`$-axis is defined from the condition $`J=\mathrm{\Phi }_{\mathrm{\Omega }_x}\widehat{J}_x\mathrm{\Phi }_{\mathrm{\Omega }_x}=\sqrt{I(I+1)}.`$ (27) where $`I`$ is total nuclear spin. $`U_k(𝒓)`$ and $`V_k(𝒓)`$ are quasiparticle Dirac spinors and $`E_k`$ denotes the quasiparticle energies. The pairing potential (field) $`\widehat{\mathrm{\Delta }}`$ in Eq. (19) is given by $`\widehat{\mathrm{\Delta }}\mathrm{\Delta }_{ab}={\displaystyle \frac{1}{2}}{\displaystyle \underset{cd}{}}V_{abcd}^{pp}\kappa _{cd}`$ (28) where the indices $`a,b,\mathrm{}`$ denote quantum numbers which specify the single-particle states with the space coordinates $`𝒓`$ as well as the Dirac and isospin indices $`s`$ and $`\tau `$. It contains the pairing tensor This quantity is sometimes called as abnormal density. $`\kappa `$ $`\kappa =V^{}U^T`$ (29) and the matrix elements $`V_{abcd}^{pp}`$ of the effective interaction in the $`pp`$-channel. The matrix elements $`V_{abcd}^{pp}`$ in the pairing channel can, in principle, be derived as a one-meson exchange interaction by eliminating the mesonic degrees of freedom in the model Lagrangian as discussed in detail in the relativistic Hartree-Bogoliubov model developed in Ref. . However, the resulting pairing matrix elements obtained with the standard sets of the RMF theory are unrealistically large. In nuclear matter the strong repulsion produced by the exchange of vector mesons at short distances results in a pairing gap at the Fermi surface that is by factor of 3 too large. In addition, standard RMF parameters do not reproduce scattering data in the $`S_0`$-channel, which is necessary for a reasonable description of pairing correlations. On the other hand, since we are using effective forces, there is no fundamental reason to have the same interaction both in the particle-hole and particle-particle channel . In a first-order approximation, the effective interaction contained in the mean field $`\widehat{\mathrm{\Gamma }}`$ is a $`G`$ matrix, i.e. the sum over all ladder diagrams. On the contrary, the effective force in the $`pp`$ channel (pairing potential $`\widehat{\mathrm{\Delta }}`$) should be the $`K`$ matrix, the sum of all diagrams irreducible in $`pp`$ direction. Although encouraging results have been reported in applications to nuclear matter , a microscopic and fully relativistic derivation of the pairing force starting from the Lagrangian of quantum hadrodynamics still cannot be applied to realistic nuclei. Thus we follow the prescription of Ref. and use a phenomenological interaction of Gogny type with finite range (see Eq. (30) below) in the particle-particle channel. Such a procedure provides both the automatic cutoff of high-momentum components and, as follows from non-relativistic and relativistic studies, a reliable description of pairing properties in finite nuclei. Although this procedure formally breaks the Lorentz structure of the RMF equations, one has to keep in mind that pairing itself is a completely non-relativistic phenomenon. Relativistic effects such as the nuclear saturation mechanism due to the cancellation between attractive scalar and repulsive vector potentials, spin-orbit splitting and the admixture of small components through the kinetic term $`i𝜶\mathbf{}`$ of the Dirac equation are only important for the mean field part of Hartree-Bogoliubov theory and have only a negligible counterpart in the pairing field. The pairing density $`\kappa `$ and the corresponding pairing field result from the scattering of pairs in the vicinity of the Fermi surface. The pairing density is concentrated in an energy window of a few MeV around the Fermi level, i.e. the contributions from the small components of the wave functions to the pairing tensor $`\kappa `$ are very small. Thus in the present version of the CRHB theory, pairing correlations are only considered between the positive energy states. Consequently, it is justified to approximate the pairing force by the best currently available non-relativistic interaction: the pairing part of the Gogny force. At present, this is certainly more realistic than the use of a one-meson exchange force in the pairing channel, since this type of interactions has never been optimized for the description of pairing properties in finite nuclei. Thus the word relativistic in CRHB applies only to the Hartree particle-hole channel of this theory. The phenomenological Gogny-type finite range interaction is given by $`V^{pp}(1,2)=f{\displaystyle \underset{i=1,2}{}}e^{[(𝒓_1𝒓_2)/\mu _i]^2}\times (W_i+B_iP^\sigma H_iP^\tau M_iP^\sigma P^\tau )`$ (30) where $`\mu _i`$, $`W_i`$, $`B_i`$, $`H_i`$ and $`M_i`$ $`(i=1,2)`$ are the parameters of the force and $`P^\sigma `$ and $`P^\tau `$ are the exchange operators for the spin and isospin variables, respectively. This interaction is density-independent. In most of the applications of the CRHB theory in the present article, the parameter set D1S (see also Table I) is employed for the Gogny force. Note also that an additional factor $`f`$ affecting the strength of the Gogny force is introduced in Eq. (30). This is motivated by the fact that previous studies of different nuclear phenomena within the Relativistic Hartree-Bogoliubov theory followed the prescription of Ref. where $`f=1.15`$ was employed. It turns out, however, that the rotational properties of the SD bands are very well described with $`f=1.0`$ if the approximate particle number projection is performed by means of the Lipkin-Nogami method. Thus the value $`f=1.0`$ is used in the following if no other value is specified. In Hartree-(Fock)-Bogoliubov calculations the size of the pairing correlations is usually measured in terms of the pairing energyThis quantity is sometimes called as particle-particle correlation energy, see for example Ref. defined as $`E_{pairing}={\displaystyle \frac{1}{2}}\text{Tr}(\mathrm{\Delta }\kappa ).`$ (31) This is not an experimentally accessible quantity, but it is a measure for the size of the pairing correlations in the theoretical calculations. An alternative way to look for the size of pairing correlations is to use the BCS-like pairing energies defined as a difference between the total energies obtained at a given spin $`I`$ in the calculations with ($`E_{\mathrm{CRHB}}(I)`$) and without ($`E_{\mathrm{CRMF}}(I)`$) pairing $`E_{\mathrm{BCS}}(I)=E_{\mathrm{CRHB}}(I)E_{\mathrm{CRMF}}(I)`$ (32) The quantities $`E_{\mathrm{BCS}}`$ and $`E_{pairing}`$ coincide only in the BCS approximation. The disadvantage of the use of the BCS-like pairing energies is related to the fact that the calculations with and without pairing should be performed in order to define this quantity in a proper way. The total energy of system in the laboratory frame is given by $`E_{\mathrm{CRHB}}=E^\mathrm{F}+E^\mathrm{B}`$ (33) where $`E^\mathrm{F}`$ and $`E^\mathrm{B}`$ are the contributions from fermionic and mesonic (bosonic) degrees of freedom. Fermionic energies $`E^\mathrm{F}`$ are given by $`E^\mathrm{F}=E_{part}+\mathrm{\Omega }_xJ+E_{pairing}+E_{cm}`$ (34) where $`E_{part}=Tr(h_D\rho ),J=Tr(j_x\rho ),`$ (35) are the particle energy and the expectation value of the total angular momentum along the rotational axis and $`E_{cm}={\displaystyle \frac{3}{4}}\mathrm{}\omega _0={\displaystyle \frac{3}{4}}41A^{1/3}\mathrm{MeV}`$ (36) is the correction for the spurious center-of-mass motion approximated by its value in a non-relativistic harmonic oscillator potential. The bosonic energies $`E^\mathrm{B}`$ in the laboratory frame are given by $`E^\mathrm{B}=`$ $`{\displaystyle \frac{1}{2}}{\displaystyle 𝑑𝒓[g_\sigma \sigma (𝒓)\rho _s(𝒓)+\frac{1}{3}g_2\sigma ^3(𝒓)+\frac{1}{2}g_3\sigma ^4(𝒓)]}`$ (42) $`{\displaystyle \frac{1}{2}}g_\omega {\displaystyle 𝑑𝒓[\omega _0(𝒓)\rho _v^{is}(𝒓)𝝎(𝒓)𝒋^{is}(𝒓)]}`$ $`{\displaystyle \frac{1}{2}}g_\rho {\displaystyle 𝑑𝒓[\rho _0(𝒓)\rho _v^{iv}(𝒓)𝝆(𝒓)𝒋^{iv}(𝒓)]}`$ $`{\displaystyle \frac{1}{2}}e{\displaystyle 𝑑𝒓\left[A_0(𝒓)\rho _v^p(𝒓)+𝑨(𝒓)𝒋^p(𝒓)\right]}`$ $`+\mathrm{\Omega }_x^2{\displaystyle }d𝒓[\sigma (𝒓)\widehat{L}_x^2\sigma (𝒓)\omega _0(𝒓)\widehat{L}_x^2\omega _0(𝒓)+𝝎(𝒓)(\widehat{L}_x+\widehat{S}_x)^2𝝎(𝒓)`$ $`\rho _0(𝒓)\widehat{L}_x^2\rho _0(𝒓)+𝝆(𝒓)(\widehat{L}_x+\widehat{S}_x)^2𝝆(𝒓)]`$ In the systems with broken time-reversal symmetry the spatial components of the $`\omega `$-mesons give larger contributions to the total energy than the ones of the $`\rho `$-mesons because of the isovector nature of the $`\rho `$-meson. One should also note that the contribution of the terms proportional to $`\mathrm{\Omega }_x^2`$ to the total energy is very small being typically in the range of $`1020`$ keV at the highest frequencies of interest in the $`A190`$ mass region and thus, in general, can be neglected. ### B The Lipkin-Nogami Method In recent years it has become clear that a proper treatment of pairing correlations is required to describe many nuclear properties. One should note that the Bogoliubov transformation is not commutable with the nucleon number operator and consequently the resulting wave function does not correspond to a system having a definite number of protons and neutrons. The best way to deal with this problem would be to perform an exact particle number projection before the variation . However, for heavy nuclei this has been realized so far only for non-relativistic separable models in part due to the fact that such calculations are expected to be extremely time-consuming for realistic interactions. As a result, an approximate particle number projection by means of the Lipkin-Nogami (LN) method (further APNP(LN)) is most widely used just because of its simplicity. As illustrated by the results of non-relativistic and recent relativistic calculations, the application of this method considerably improves an agreement with experiment especially for rotational properties of nuclei compared with the calculations without particle number projection. Our derivation of the Lipkin-Nogami method is based on the Kamlah expansion which is an approximation to the exact projection methods. It is applied to the two-body interaction $`V^{pp}`$ in the pairing channel. Note that the Gogny force in the pairing channel is density independent in the CRHB theory. For the derivation of the Lipkin-Nogami method with density dependent forces see Refs. . In the following discussion, we will limit ourselves to one type of particles. In the numerical calculations we use this approach for protons and neutrons separately. The eigenstates $`|\mathrm{\Psi }^N`$ of the particle number operator $`\widehat{N}`$ can be constructed from particle number violating Bogoliubov wave functions $`|\mathrm{\Phi }`$ by means of the particle number projection operator $`\widehat{P}^N`$ $`|\mathrm{\Psi }^N=\widehat{P}^N|\mathrm{\Phi }={\displaystyle \frac{1}{2\pi }}{\displaystyle _0^{2\pi }}𝑑\varphi e^{i(\widehat{N}N)\varphi }|\mathrm{\Phi }`$ (43) where $`N`$ is actual particle number. With this wave function one obtains the total particle number projected energy $`E_{proj}^N={\displaystyle \frac{\mathrm{\Phi }|\widehat{H}\widehat{P}^N|\mathrm{\Phi }}{\mathrm{\Phi }|\widehat{P}^N|\mathrm{\Phi }}}={\displaystyle \frac{_0^{2\pi }𝑑\varphi e^{i\varphi N}h(\varphi )}{_0^{2\pi }𝑑\varphi e^{i\varphi N}n(\varphi )}}.`$ (44) In the case of large particle number and strong deformations in gauge space one expects that the Hamiltonian and norm overlap integrals $`h(\varphi )=\mathrm{\Phi }|\widehat{H}e^{i\varphi \widehat{N}}|\mathrm{\Phi }\text{ and }n(\varphi )=\mathrm{\Phi }|e^{i\varphi \widehat{N}}|\mathrm{\Phi },`$ (45) are sharply peaked at $`\varphi =0`$ and are very small elsewhere. In addition, the quotient $`h(\varphi )/n(\varphi )`$ should be a rather smooth function. In such a case one can make an expansion of $`h(\varphi )`$ in terms of $`n(\varphi )`$ in the following way $`h(\varphi )={\displaystyle \underset{m=0}{\overset{M}{}}}\lambda _m\widehat{𝒩}^mn(\varphi )`$ (46) where the Kamlah operator $`\widehat{𝒩}={\displaystyle \frac{1}{i}}{\displaystyle \frac{}{\varphi }}\mathrm{\Phi }|\widehat{N}|\mathrm{\Phi }`$ (47) represents the particle number operator in the space of the gauge angle $`\varphi `$ and $`\lambda _m`$ are constants. Mathematically, Eq. (46) essentially corresponds to a Taylor expansion of the Fourier transformed function $`h(\varphi )/n(\varphi )`$ . $`M`$ in Eq. (46) represents the order of expansion, for $`M\mathrm{}`$ this equation is exact. Since we need $`h(\varphi )`$ only in the vicinity of $`\varphi =0`$ we can already hope to get a very good approximation for this function with a few non-vanishing constants $`\lambda _0,\lambda _1,\lambda _2,\mathrm{}`$, if they are properly adjusted. The number $`M`$ certainly depends on the widths of the overlap integral $`n(\varphi )`$ which reflects the symmetry violation. The larger the symmetry violation, the better the approximation will already be for low $`M`$-values. Inserting Kamlah operator into Eq. (46) one gets $`h(\varphi )={\displaystyle \underset{m=0}{\overset{M}{}}}\lambda _m\mathrm{\Phi }|(\mathrm{\Delta }\widehat{N})^me^{i\varphi \widehat{N}}|\mathrm{\Phi }\text{ with }\mathrm{\Delta }\widehat{N}=\widehat{N}\mathrm{\Phi }|\widehat{N}|\mathrm{\Phi }.`$ (48) The expansion coefficients $`\lambda _m`$ are determined by applying the operators $`\widehat{𝒩}^k`$ $`(k=0,1,\mathrm{}M)`$ on Eq. (46) $`\widehat{𝒩}^kh(\varphi )={\displaystyle \underset{m=0}{\overset{M}{}}}\lambda _m\widehat{𝒩}^{m+k}n(\varphi ),\text{ }k=0\mathrm{}M.`$ (49) This yields $`\mathrm{\Phi }|(\mathrm{\Delta }\widehat{N})^k\widehat{H}e^{i\varphi \widehat{N}}|\mathrm{\Phi }={\displaystyle \underset{m=0}{\overset{M}{}}}\lambda _m\mathrm{\Phi }|(\mathrm{\Delta }\widehat{N})^{m+k}e^{i\varphi \widehat{N}}|\mathrm{\Phi }.`$ (50) Taking the limit $`\varphi 0`$ this results in a system of the $`M+1`$ linear equations $`\mathrm{\Phi }|[\widehat{H}{\displaystyle \underset{m=0}{\overset{M}{}}}\lambda _m(\mathrm{\Delta }\widehat{N})^m](\mathrm{\Delta }\widehat{N})^k|\mathrm{\Phi }=0`$ (51) for the constants $`\lambda _m`$. The case of $`M=1`$ corresponds to CRHB theory discussed in Sect. II A. In the following we will use the shorthand notation $`\widehat{O}=\mathrm{\Phi }|\widehat{O}|\mathrm{\Phi }`$. In the case of $`M=2`$, the solution of the system of Eqs. (51) under the particle number constraint $`\mathrm{\Delta }\widehat{N}=0`$ leads to the following values of $`\lambda _0,\lambda _1,\lambda _2`$: $`\lambda _0`$ $`=`$ $`\widehat{H}\lambda _2(\mathrm{\Delta }\widehat{N})^2,`$ (52) $`\lambda _1`$ $`=`$ $`{\displaystyle \frac{\widehat{H}\mathrm{\Delta }\widehat{N}\lambda _2(\mathrm{\Delta }\widehat{N})^3}{(\mathrm{\Delta }\widehat{N})^2}},`$ (53) $`\lambda _2`$ $`=`$ $`{\displaystyle \frac{\widehat{H}[(\mathrm{\Delta }\widehat{N})^2(\mathrm{\Delta }\widehat{N})^2]\widehat{H}\mathrm{\Delta }\widehat{N}(\mathrm{\Delta }\widehat{N})^3/(\mathrm{\Delta }\widehat{N})^2}{(\mathrm{\Delta }\widehat{N})^4(\mathrm{\Delta }\widehat{N})^2^2(\mathrm{\Delta }\widehat{N})^3^2/(\mathrm{\Delta }\widehat{N})^2}},`$ (54) where the moments of the operator $`\mathrm{\Delta }\widehat{N}`$ are given by $`(\mathrm{\Delta }\widehat{N})^2`$ $`=`$ $`2Tr[\chi ],`$ (55) $`(\mathrm{\Delta }\widehat{N})^3`$ $`=`$ $`4Tr[\gamma \chi ],`$ (56) $`(\mathrm{\Delta }\widehat{N})^4`$ $`=`$ $`(\mathrm{\Delta }\widehat{N})^2+8Tr[\chi (16\chi )],`$ (57) with $`\chi =\rho (1\rho )\mathrm{and}\gamma =12\rho .`$ (58) Because $`\widehat{𝒩}`$ is a hermitian operator with respect to the integration on $`\varphi `$ between 0 and $`2\pi `$, one gets a very simple result for the particle number projected energy to order $`M`$ $$E_{proj}^N=\underset{m=0}{\overset{M}{}}\lambda _m(N\widehat{N})^m.$$ (59) In a full variation after projection method one should vary $`E_{proj}^N`$. The Lipkin-Nogami method consists in treating $`\lambda _2`$ as a constant during the variation, with its value defined according to Eq. (54) being readjusted self-consistently at each iteration. This leads to the variational equation $`{\displaystyle \frac{\delta }{\delta \mathrm{\Phi }}}\mathrm{\Phi }|\widehat{H}\lambda _2(\mathrm{\Delta }\widehat{N})^2|\mathrm{\Phi }\lambda {\displaystyle \frac{\delta }{\delta \mathrm{\Phi }}}\mathrm{\Phi }|\widehat{N}|\mathrm{\Phi }=0,`$ (60) which means that we have to minimize the expectation value of the particle number projected energy within the set of product wave functions $`|\mathrm{\Phi }`$ under subsidiary condition that $`\lambda `$ is determined by the particle number constraint $`\mathrm{\Phi }|\widehat{N}|\mathrm{\Phi }=N,`$ (61) provided that the condition $`\lambda =\lambda _1`$ is fulfilled and the terms proportional to $`\frac{\delta }{\delta \mathrm{\Phi }}\lambda _2`$ are neglected. This means in particular that the Lipkin-Nogami method violates the variational principle. An extension of the Lipkin-Nogami method to a full variation of the Kamlah expansion of the projected energy up to second order is very complicated. So far it has only been carried out in non-relativistic theories . The formulation of the Lipkin-Nogami method presented above is given for one kind of nucleons (protons or neutrons). In realistic calculations we perform it simultaneously for protons and neutrons. It can easily be shown that for $`M=2`$ there is no coupling between protons and neutrons. The evaluation of the term $`\lambda _2`$ for a general two-particle interaction $`V^{pp}`$ is given as $`\lambda _2={\displaystyle \frac{1}{4}}{\displaystyle \frac{2Tr_1Tr_1(\kappa \kappa ^+\overline{v}\kappa \kappa ^+)Tr_2Tr_2(\kappa ^{}\rho \overline{v}\sigma \kappa )}{[Tr(\kappa \kappa ^+)]^22Tr(\kappa \kappa ^+\kappa \kappa ^+)}},`$ (62) where $`\sigma =1\rho `$ and $`\overline{v}_{abcd}=ab|V^{pp}|cddc`$ is antisymmetrized matrix element of the two-particle interaction $`V^{pp}`$. The trace $`Tr_1`$ represents the summation in the particle-hole channel, while $`Tr_2`$ in particle-particle channel . In Relativistic Hartree-Bogoliubov theory with the Gogny force in the pairing channel, the sum over the particle-hole part is zero, since according to our definition the two-particle interaction $`V^{pp}`$ acts only between the fermions. In the particle-hole channel we have only classical fields conserving the particle number. As a result, the $`\lambda _2`$ value used in the CRHB calculations with APNP(LN) is given by $`\lambda _2={\displaystyle \frac{1}{4}}{\displaystyle \frac{Tr_2Tr_2(\kappa ^{}\rho \overline{v}\sigma \kappa )}{[Tr(\kappa \kappa ^+)]^22Tr(\kappa \kappa ^+\kappa \kappa ^+)}}.`$ (63) Note that in the harmonic oscillator basis used presently in the CRHB(+LN) calculations, the density matrix $`\rho `$ and pairing tensor $`\kappa `$ entering into Eq. (63) are real. In Ref. it is shown in detail that Eq. (63) can be represented by $`\lambda _2={\displaystyle \frac{1}{4}}{\displaystyle \frac{E_p[\gamma \kappa ]E_p[\kappa ]}{[Tr(\kappa \kappa ^+)]^22Tr(\kappa \kappa ^+\kappa \kappa ^+)}},`$ (64) where the unprojected pairing energy functional is given by $`E_p[\kappa ]={\displaystyle \frac{1}{4}}Tr_2Tr_2(\kappa ^{}\overline{v}\kappa ).`$ (65) Since the modified pairing tensor $`\gamma \kappa `$ is much smaller than the pairing tensor $`\kappa `$, the main contribution to $`\lambda _2`$ comes from the pairing energy. The application of the Lipkin-Nogami method leads to a modification of the Hartree-Bogoliubov equations for the fermions, while the mesonic part of the CRHB theory is not affected. This modification is obtained by the restricted variation of $`\lambda _2(\mathrm{\Delta }N)^2`$, namely, $`\lambda _2`$ is not varied and its value is calculated self-consistently using Eq. (64) in each step of the iteration. One should note that the form of the CRHB+LN equations is not unique (see Refs. for details). In the general case, the CRHF+LN equation contains a parameter $`\eta (\eta =0,\pm 1)`$ and is given by $`\left(\begin{array}{cc}\widehat{h}_D(\eta )\lambda (\eta )\mathrm{\Omega }_x\widehat{J_x}& \widehat{\mathrm{\Delta }}(\eta )\\ \widehat{\mathrm{\Delta }}^{}(\eta )& \widehat{h}_D^{}(\eta )+\lambda (\eta )\mathrm{\Omega }_x\widehat{J_x}^{}\end{array}\right)\left(\begin{array}{c}U(𝒓)\\ V(𝒓)\end{array}\right)_k=E_k(\eta )\left(\begin{array}{c}U(𝒓)\\ V(𝒓)\end{array}\right)_k`$ (66) where $`\widehat{h}_D(\eta )`$ $`=`$ $`\widehat{h}_D+2\lambda _2[(1+\eta )\rho Tr(\rho )],`$ (67) $`\widehat{\mathrm{\Delta }}(\eta )`$ $`=`$ $`\widehat{\mathrm{\Delta }}2\lambda _2(1\eta )\kappa ,`$ (68) $`\lambda (\eta )`$ $`=`$ $`\lambda _1+\lambda _2[1+\eta ],`$ (69) $`E_k(\eta )`$ $`=`$ $`E_k\eta \lambda _2.`$ (70) With these definitions and neglecting for simplicity $`2\lambda _2Tr(\rho )`$ in $`\widehat{h}_D(\eta )`$ in the following discussion it is clear that the case of $`\eta =+1`$ corresponds to the shift of whole modification into the particle-hole channel of the CRHB+LN theory: $`\widehat{h}_D\widehat{h}_D+4\lambda _2\rho `$ leaving pairing potential $`\widehat{\mathrm{\Delta }}`$ unchanged. The case of $`\eta =1`$ correspond to the shift of the modification into the particle-particle channel $`\widehat{\mathrm{\Delta }}\widehat{\mathrm{\Delta }}4\lambda _2\kappa `$ leaving $`\widehat{h}_D`$ unchanged. An intermediate situation is obtained in the case of $`\eta =0`$: $`\widehat{h}_D\widehat{h}_D+2\lambda _2\rho `$, $`\widehat{\mathrm{\Delta }}\widehat{\mathrm{\Delta }}2\lambda _2\kappa `$. One must note that the eigenvalues $`E_k(\eta )`$ of the CRHB+LN equations are identical to the quasiparticle energies $`E_k`$ only in the case of $`\eta =0`$. In the present calculations we are using the case of $`\eta =+1`$ which provides reasonable numerical stability of the CRHB+LN equations. ### C Physical observables and details of the calculations Because, with a few exceptions, the observed SD bands are not linked to the low-spin level schemes, the absolute angular momentum quantum numbers $`I`$ are not known experimentally. Thus the dynamic moment of inertia $`J^{(2)}`$ which contains only the differences $`\mathrm{\Delta }I=2`$ plays an important role in our understanding of their structure. In the calculations, the rotational frequency $`\mathrm{\Omega }_x`$, the kinematic moment of inertia $`J^{(1)}`$ and the dynamic moment of inertia $`J^{(2)}`$ are defined by $`\mathrm{\Omega }_x={\displaystyle \frac{dE}{dJ}},J^{(1)}(\mathrm{\Omega }_x)=J\left\{{\displaystyle \frac{dE}{dJ}}\right\}^1={\displaystyle \frac{J}{\mathrm{\Omega }_x}},J^{(2)}(\mathrm{\Omega }_x)=\left\{{\displaystyle \frac{d^2E}{dJ^2}}\right\}^1={\displaystyle \frac{dJ}{d\mathrm{\Omega }_x}}.`$ (71) Experimental quantities such as a the rotational frequency, the kinematic and dynamic moments of inertia are extracted from the observed energies of $`\gamma `$-transitions within a band according to the prescription given in Sect. 4.1 of Ref. . One should note that the kinematic moment of inertia depends on the absolute values of the spin which in most cases are not know in SD bands. The charge quadrupole $`Q_0`$ and mass hexadecupole $`Q_{40}`$ moments are calculated by using the expressions $`Q_0`$ $`=`$ $`e\sqrt{{\displaystyle \frac{16\pi }{5}}}\sqrt{r^2Y_{20}_p^2+2r^2Y_{22}_p^2}`$ (72) $`Q_{40}`$ $`=`$ $`r^4Y_{40}_p+r^4Y_{40}_n`$ (73) where the labels $`p`$ and $`n`$ are used for protons and neutrons, respectively and $`e`$ is the electrical charge. The transition quadrupole moment $`Q_t`$ for a triaxially deformed nucleus is calculated by the following expression $`Q_t=e\sqrt{{\displaystyle \frac{16\pi }{5}}}r^2Y_{20}_p{\displaystyle \frac{\mathrm{cos}(\gamma +30^{})}{\mathrm{cos}(30^{})}}`$ (74) where the $`\gamma `$-deformation (of the proton subsystem) is defined by $`\mathrm{tan}\gamma ={\displaystyle \frac{r^2Y_{22}_p}{r^2Y_{20}_p}}`$ (75) In the limit $`\gamma 0`$ which is typical for SD bands in the $`A190`$ region, $`Q_t`$ is almost identical to $`Q_0`$. The comparison between calculated transition quadrupole moments and available experimental data has been performed in Ref. and thus it will not be repeated here. They agree with each other if the uncertainties due to stopping powers are taken into account in the experimental data. In addition, it was concluded that more accurate and consistent experimental data on $`Q_t`$ is needed in order to carry out such a comparison in more detail. In the absence of experimentally known spins, an effective alignment approach plays an extremely important role in the definition of the structure and relative spins of SD bands. The effective alignment of two bands (A and B) is simply the difference between their spins at constant rotational frequency $`\mathrm{\Omega }_x`$ : $`i_{eff}^{\mathrm{B},\mathrm{A}}(\mathrm{\Omega }_x)=I^\mathrm{B}(\mathrm{\Omega }_x)I^\mathrm{A}(\mathrm{\Omega }_x)`$ (76) Experimentally, $`i_{eff}`$ includes the effects associated with the change of the number of the particles and the relevant changes in the alignments of the single-particle orbitals, in the deformation, in pairing etc. between two bands. This physical observable has been used frequently in unpaired calculations for the configuration and spin assignments of SD bands in the $`A60`$ and $`A140150`$ mass regions (see Refs. and references therein) and for the study of relative properties of smooth terminating bands in the $`A110`$ mass region . The effective alignment approach exploits the fact that the spin is quantized, integer for even nuclei and half-integer for odd nuclei and that it is furthermore constrained by signature. One should note that with the configurations and specifically the signatures fixed, the relative spins of observed bands can only change in steps of $`\pm 2\mathrm{}n`$, where $`n`$ is an integer number. The excitation energies of SD bands relative to the ground state are known definitely in <sup>194</sup>Hg and <sup>194</sup>Pb and tentatively in <sup>192</sup>Pb. The RMF theory with a somewhat simplistic pairing and no particle number projection excellently reproduces these data . It would be interesting to see how this result will be changed when the Gogny force is used in the pairing channel and APNP(LN) is performed. Considering, however, that such an investigation will require a constraint on the quadrupole moment and thus will be extremely time consuming within a three-dimensional CRHB computer code, we will leave this question open for future studies. The CRHB-equations are solved in the basis of an anisotropic three-dimensional harmonic oscillator in Cartesian coordinates. The same basis deformation $`\beta _0=0.5`$, $`\gamma =0^{}`$ and oscillator frequency $`\mathrm{}\omega _0=41`$A<sup>-1/3</sup> MeV have been used for all nuclei. All fermionic and bosonic states belonging to the shells up to $`N_F=14`$ and $`N_B=16`$ are taken into account in the diagonalization of the Dirac equation and the matrix inversion of the Klein-Gordon equations, respectively. The detailed investigations performed for <sup>194</sup>Hg indicate that this truncation scheme provides reasonable numerical accuracy. The values of the kinematic moment of inertia $`J^{(1)}`$ and the transition quadrupole moment $`Q_t`$ obtained with such a truncation of the basis are different from the ones obtained in the fermionic basis with $`N_F=17`$ by less than 1%. The accuracy of the calculated mass hexadecupole moment $`Q_{40}`$ is a bit lower being $`1.5`$%. The numerical accuracy of the calculations of the physical observables within the employed basis has also been tested using different combinations of $`\beta _0`$ and $`\mathrm{}\omega _0`$, namely $`\beta _0=0.4`$, 0.5 and 0.6 as well as $`\mathrm{}\omega _0=38A^{1/3}`$, $`41A^{1/3}`$, $`44A^{1/3}`$, $`47A^{1/3}`$, $`50A^{1/3}`$, $`53A^{1/3}`$ MeV. A similar level of accuracy has been obtained for $`J^{(1)}`$, but the new estimations of accuracy for the calculation of $`Q_t`$ and $`Q_{40}`$ are around 2% and 3%, respectively. The numerical errors for the total energy (in %) are even smaller. At $`\mathrm{\Omega }_x=0.0`$ MeV, the single-particle orbitals are labeled by means of the asymptotic quantum numbers $`[Nn_z\mathrm{\Lambda }]\mathrm{\Omega }`$ (Nilsson quantum numbers) of the dominant component of the wave function and superscripts to the orbital labels (e.g. $`[651]1/2^+`$) are used to indicate the sign of the signature $`r`$ for that orbital $`(r=\pm i)`$. ## III The impact of pairing and particle number projection In the present section, the impact of pairing and approximate particle number projection will be studied in detail on the example of the yrast SD bands observed in <sup>192</sup>Hg and <sup>194</sup>Pb. In order to differentiate different types of calculations, the following abbreviations are introduced: * CRMF - cranked relativistic mean field calculations without pairing correlations, * CRHB - cranked relativistic Hartree+Bogoliubov calculations without particle number projection, * CRHB+LN - cranked relativistic Hartree+Bogoliubov calculations with approximate particle number projection by means of the Lipkin-Nogami method (APNP(LN)). In the following the lines representing the results of different CRHB(+LN) calculations will be labelled in the figures by $`\mathrm{𝐑𝐌𝐅𝐬𝐞𝐭}+f\mathrm{𝐆𝐨𝐠𝐬𝐞𝐭}+\mathrm{𝐋𝐍}`$ (77) where RMFset and Gogset are the RMF and Gogny forces used in the calculations, $`f`$ is the scaling factor for the Gogny force (see Eq. (30)). Note that $`f`$ will in general be omitted when $`f=1.0`$. LN is used only in the cases when APNP(LN) is performed in the calculations. The lines showing the results of the CRMF calculations will be labelled by RMFset \[CRMF\]. We start from the results of the CRMF calculations. Neutron and proton single-routhian diagrams calculated with the parameter set NL1 (see Table II) and drawn along the deformation path of the yrast SD configurations in these two nuclei are shown in Fig. 1. The shell structure of <sup>192</sup>Hg is dominated by the large $`Z=80`$ and $`N=112`$ SD shell gaps. The same shell gaps appear also in the calculations with the set NL3, see Fig. 2, but compared with the set NL1 the $`Z=80`$ SD shell gap is somewhat smaller while the $`N=112`$ SD shell gap is more pronounced. Similar shell gaps are seen also in <sup>194</sup>Pb (see right panels of Fig. 1). Considering that the $`Z=82`$ SD shell gap decreases with increasing rotational frequency due to down-sloping $`\pi [651]1/2^+`$ orbital, it is more reasonable to consider the nucleus <sup>192</sup>Hg as a doubly magic in this mass region. While the NL1 and NL3 forces give the same set of single-particle orbitals in the vicinity of magic SD shell gaps, their relative ordering (energies) is somewhat different. As studied in detail in Ref. (Section 6.4), the difference in the single-particle energies at superdeformation when different RMF parametrizations are used is to a great extend connected with their energy differences at spherical shape. Note that the proton single-particle spectra in the vicinity of the $`Z=80`$ SD shell gap in the $`A190`$ mass region reveal large similarities with the neutron single-particle spectra in the vicinity of the $`N=80`$ SD shell gap in the $`A140150`$ mass region, see Fig. 2 in the present article and Fig. 17 in Ref. . Figs. 1 and 2 show some similarities and differences with the results obtained in the non-relativistic calculations with the Wood-Saxon potential (see Fig. 3 in Ref. ), Skyrme forces (see Figs. 6 and 7 in Ref. ) and Gogny forces (see Fig. 5 in Ref. ). Both in relativistic and non-relativistic calculations a similar shell structure and a similar set of single-particle states (although their relative energies are different) appear in the vicinity of the SD shell gaps. The main difference with non-relativistic calculations is related to the larger size of the SD shell gaps and to the lower level density in the vicinity of these gaps. These features are connected with low effective mass of RMF theory, see the discussion in Ref. for details. As can be seen in Fig. 3, the CRMF calculations do not reproduce the experimental kinematic and dynamic moments of inertia. In the case of <sup>192</sup>Hg, the calculated values of $`J^{(1)}`$ and $`J^{(2)}`$ are almost constant, while in the case of <sup>194</sup>Pb the unpaired proton band crossing is clearly seen. It originates from the interaction between $`\pi [642]5/2^+`$ and $`\pi [651]1/2^+`$ orbitals, see bottom panels in Fig. 1. The orbital $`\pi [642]5/2^+`$ is occupied before band crossing in the lowest SD configuration, while the orbital $`\pi [651]1/2^+`$ is occupied after band crossing. Note that in the $`A150`$ mass region, the crossings observed in some bands of the nuclei around <sup>147</sup>Gd have been attributed to the unpaired interaction of these two orbitals, but in the neutron subsystem, see Refs. . The inclusion of pairing without particle number projection somewhat improves the agreement with experiment, see Fig. 3. Indeed, the slope of the calculated kinematic and dynamic moments of inertia at low rotational frequencies is coming closer to experiment, but still the disagreement is considerable. In addition, a proton pairing collapse takes place in the CRHB calculations. This collapse is calculated in <sup>194</sup>Pb at $`\mathrm{\Omega }_x0.35`$ MeV and in <sup>192</sup>Hg at $`\mathrm{\Omega }_x0.10`$ MeV, see Figs. 4a and 4b. In the case of <sup>194</sup>Pb it is correlated with the alignment of the $`\pi [651]1/2^+`$ orbital which reveals itself in a sharp increase of $`J^{(2)}`$, see Fig. 3a. This crossing coincides with the one seen in the CRMF calculations. At higher frequencies, the calculated kinematic and dynamic moments of inertia for the proton subsystem and the transition quadrupole moments obtained in the CRMF and CRHB calculations are very close, see Figs. 3 and 4c,d, respectively. However, at lower frequencies proton pairing is still important and the results of the calculations with and without pairing for proton $`J^{(1)}`$ and $`J^{(2)}`$ values are different. Contrary to the proton subsystem, the pairing correlations in the neutron subsystem do not collapse in the rotational frequency range under consideration. Neutron pairing energies $`E_{pairing}^\nu `$, being larger in absolute value than the proton ones at $`\mathrm{\Omega }_x=0.0`$ MeV, start around $`6`$ MeV and then smoothly decrease in absolute value with increasing rotational frequency coming close to 0 MeV at $`\mathrm{\Omega }_x=0.5`$ MeV, see Figs. 4a,b. At low and medium rotational frequencies, there is a considerable difference between the corresponding neutron moments of inertia ($`J^{(1)}`$ or $`J^{(2)}`$) obtained in the CRMF and CRHB calculations, see Fig. 3. This difference, however, is small at the highest rotational frequencies, where the magnitude of neutron pairing correlations is negligible. Note that neutron pairing correlations alone do not lead to a considerable modification of transition quadrupole moments $`Q_t`$ compared with the CRMF calculations, see an example of <sup>192</sup>Hg (Figs. 4b,d). On the contrary, in <sup>194</sup>Pb, where both neutron and proton pairing persist up to $`\mathrm{\Omega }_x0.35`$ MeV, the transition quadrupole moment $`Q_t`$ obtained in the CRHB calculations is larger than the one of the CRMF calculations. The inclusion of approximate particle number projection by means of the Lipkin-Nogami method leads to a very good agreement between the calculated and experimental kinematic and dynamic moments of inertia, see Fig. 3. The correlations induced by the Lipkin-Nogami prescription compared with CRHB produce the desired effects, namely, to diminish the kinematic moments of inertia at all frequencies, to diminish the dynamic moments of inertia at low and medium rotational frequencies and to delay proton and neutron alignments to higher frequencies. These effects are due to stronger pairing correlations seen in CRHB+LN compared with CRHB (Figs. 4a,b). Contrary to the CRHB approach where the proton pairing collapses at some frequencies, no such collapse appears in the CRHB+LN calculations in the rotational frequency range under consideration. Thus the proton and neutron subsystems remain correlated even at $`\mathrm{\Omega }_x=0.5`$ MeV. The decrease of the magnitude of pairing correlations with increasing rotational frequency seen both in CRHB and CRHB+LN is due to a strong Coriolis anti-pairing effect. The transition quadrupole moments $`Q_t`$ calculated in CRHB+LN show a different behavior as a function of rotational frequency compared with CRMF and CRHB, see Figs. 4c,d. While in the latter approaches, the $`Q_t`$ values decrease with increasing rotational frequency, they initially increase and then decrease in the CRHB+LN approach. The calculated decrease of $`Q_t`$ is due to an anti-stretching effect caused by the Coriolis force, but the differences in the behavior of $`Q_t`$ as a function of $`\mathrm{\Omega }_x`$ should be attributed to different strengths of the pairing correlations in the CRHB and CRHB+LN calculations. The increase of the kinematic and dynamic moments of inertia is a complex effect which predominantly includes the gradual alignments of the pairs of $`j_{15/2}`$ neutrons and $`i_{13/2}`$ protons and the decreasing pairing correlations with increasing rotational frequency. The quasiparticle routhian diagrams of Fig. 5 shows the typical quasiparticle spectra obtained in the CRHB+LN calculations and the alignments in the above discussed pairs. Note that the alignment of the proton pair takes place at a higher rotational frequency than the one of the neutron pair. It is reasonable to expect that the behavior of the total dynamic moment of inertia will sensitively depend on the balance of the alignments in the proton and neutron subsystems. Indeed, the total $`J^{(2)}`$ in <sup>194</sup>Pb does not show the decrease above the frequency of full alignment of the neutron pair, while such a decrease is seen in <sup>194</sup>Hg (see Figs. 3a,c). Total (proton + neutron) BCS-like pairing energies $`E_{\mathrm{BCS}}`$ (see Eq. (32)) obtained for the lowest SD configurations in <sup>192</sup>Hg and <sup>194</sup>Pb in the calculations with and without APNP(LN) are shown in Fig. 6. These quantities have more physical content than $`E_{pairing}`$ defined in Eq. (31) since they directly show the gain in binding energy due to the pairing correlations. The comparison with Figs. 4a,b allows to conclude that $`E_{\mathrm{BCS}}`$ is smaller by roughly an order of magnitude than $`E_{pairing}^{tot}=E_{pairing}^\nu +E_{pairing}^\pi `$. The CRHB calculations show rather small BCS-like pairing energies $`E_{\mathrm{BCS}}`$ which slowly converge to zero with increasing spin and almost vanish already at $`I3035\mathrm{}`$. Similar to $`E_{pairing}`$, APNP(LN) significantly (by factor $`510`$) increases the size of the BCS-like pairing energies. Although these energies decrease with increasing spin reflecting the quenching of the pairing correlations, they do not vanish even at highest calculated spins. The effective alignments $`i_{eff}`$ between the lowest SD configurations in <sup>192</sup>Hg and <sup>194</sup>Pb calculated in CRMF, CRHB and CRHB+LN are shown in Fig. 7a and are compared with experiment. It is clearly seen that the CRMF and especially the CRHB results deviate considerably from experiment in absolute value. In addition, the considerable change of the slope of $`i_{eff}`$ obtained in these calculations at $`\mathrm{\Omega }_x0.30.35`$ MeV, which is due to proton band crossing, contradicts to experimental data. Similar to the moments of inertia, APNP(LN) considerably improves the agreement between calculations and experiment also for the effective alignments. Although the values $`i_{eff}`$ calculated in CRHB+LN deviate by $`0.40.5\mathrm{}`$ from experiment, this deviation should not be considered as crucial since the compared bands differ by 2 protons. Moreover, the slope of $`i_{eff}`$ as a function of $`\mathrm{\Omega }_x`$ is rather well reproduced in the calculations. ## IV The dependence of the results on the parametrization of the mean field and pairing force. ### A The dependence of the results on the parametrization of the RMF Lagrangian. The pairing correlations depend not only on the properties of the effective pairing force, but also on the single-particle level density. A full relativistic Hartree-Bogoliubov calculation is therefore only meaningful if the Hartree field yields a reasonable single-particle spectrum . In order to investigate the dependence of the results on the parametrization of the RMF Lagrangian, the CRHB+LN calculations have been performed with the NL3 force for the lowest SD configurations in <sup>194</sup>Pb and <sup>194</sup>Hg, see Figs. 8 and 9. In addition, the NLSH force has been employed in <sup>194</sup>Hg, but due to slow convergence it was used only in a short frequency range. In all these calculations, the D1S set has been used for the Gogny force. At low rotational frequencies, total, neutron and proton kinematic and dynamic moments of inertia obtained in the calculations with NL3 are smaller than the corresponding quantities calculated with NL1, see Fig. 8. However, the increase of $`J^{(2)}`$ and $`J^{(1)}`$ as a function of rotational frequency is larger in the calculations with NL3 compared with the ones employing NL1. Thus at some frequencies, they reach each other. At even higher frequencies, the $`J^{(2)}`$ and $`J^{(1)}`$ values calculated with NL3 become larger than the ones calculated with NL1. It is also clear that the NL1 force provides better agreement with experimental kinematic and dynamic moments of inertia than NL3. The results of the calculations with NLSH (see Fig. 8d) are in even larger disagreement with experiment than the ones with NL3. One should note however that the NL3 force provides better reproduction of effective alignment in the <sup>194</sup>Hg/<sup>194</sup>Pb pair compared with the NL1 force, see Fig. 7b. In the NL1 parametrization, the $`Q_t`$ values are nearly constant as a function of $`\mathrm{\Omega }_x`$ both in <sup>194</sup>Hg and <sup>194</sup>Pb nuclei. The $`Q_t`$(<sup>194</sup>Pb) is approximately 2$`e`$b larger than $`Q_t`$(<sup>194</sup>Hg). The situation is different when NL3 force is used in the calculations. At $`\mathrm{\Omega }_x0`$ MeV, $`Q_t`$(<sup>194</sup>Pb) is approximately equal to $`Q_t`$(<sup>194</sup>Hg). This fact is most likely related to the $`N=112`$ SD shell gap which is more pronounced in the NL3 parametrization (see Fig. 2 and discussion in Sect. III). However, with increasing rotational frequency $`Q_t`$(<sup>194</sup>Pb) increases considerably with a maximum gain of $`1.5`$ $`e`$b at $`\mathrm{\Omega }_x0.35`$ MeV (Fig. 9c). On the contrary, with exception of the band crossing region the evolution of $`Q_t`$(<sup>194</sup>Hg) as a function of $`\mathrm{\Omega }_x`$ is similar to the one seen in the calculations with NL1. Comparing different parametrizations of the RMF Lagrangian, it is clear that NL1 produces the largest values of $`Q_t`$, while NLSH the smallest (Fig. 9d)). This tendency has already been seen in the $`A60`$ and $`A140150`$ regions of superdeformation . In the calculations with NL3, proton and neutron pairing energies are similar at low rotational frequencies (see Figs. 9a,b). At $`\mathrm{\Omega }_x0.15`$ MeV, proton pairing energies are larger than neutron ones. On the other hand, in the calculations with NL1 neutron pairing energies are larger than proton ones at all rotational frequencies. The comparison of single-particle energies at $`\mathrm{\Omega }_x=0.0`$ MeV obtained with the NL1 and NL3 parametrizations of the RMF theory (see Fig. 2) suggests that this is due to the larger $`N=112`$ and the smaller $`Z=80`$ SD shell gaps in the calculations with NL3. This leads to an additional quenching of neutron pairing correlations and to an increase of proton pairing correlations as compared with the case of the NL1 force. ### B The dependence of the results on the parametrization of the Gogny force. In the present section, we will study how the results of the calculations depend on the parametrization and the strength of the Gogny force. In all calculations given in this section, the set NL1 is used for the RMF Lagrangian and approximate particle number projection is performed by means of the Lipkin-Nogami method. Such a study is motivated by the fact that a precise quantitative information on the pairing correlations is not easy to extract in nuclei. There are no simple physical processes allowing one to isolate completely the pairing effects from the rest and to use them for the fit of the interaction in the particle-particle channel. Different sets of the Gogny force have been obtained by the fit to the properties of finite nuclei. As a result, these fits are more sensitive to the properties in the particle-hole channel than in the particle-particle channel since the pairing energies represent only a small portion of the total binding energies. In addition, apriori it is not clear that existing parametrizations of the Gogny force should be reliable in conjuction with the RMF theory. One should note that the moments of inertia of rotating nuclei are very sensitive to the properties of the pairing interaction . Considering that the rotation-vibration coupling is small in strongly deformed nuclei , one can try to use this fact for the definition of the best parametrization of the Gogny force which has to be used in the particle-particle channel in conjuction with the RMF theory. The parameter set D1 has been defined in Ref. based on the features of the <sup>16</sup>O, <sup>48</sup>Ca, <sup>90</sup>Zr nuclei and the pairing properties of the Sn isotopes. It turns out that in non-relativistic calculations with the Gogny force this set overestimates the strength of the pairing correlations. The D1’ set differs from the D1 set only in the strength of the spin-orbit interaction and thus it will not be employed in the present calculations since only the central force with two-finite range Gaussian terms (see Eq. (30)) of the original Gogny force is used in the CRHB theory. The D1S set differs, as follows from non-relativistic studies, from D1 by improved surface properties and by producing smaller pairing correlations. It is used in most of the calculations in the present manuscript. Recently another set of the Gogny force has been suggested in Ref. . So far it has not been applied for detailed studies of finite nuclei in non-relativistic approaches. However, comparing it with D1 and D1S (see Table I), one can conclude that it has larger similarities with D1 than with D1S. The dynamic and kinematic moments of inertia of the lowest SD configuration in <sup>192</sup>Hg calculated with D1, D1S and D1P sets are compared with experiment in Fig. 10. One can see that only the set D1S provides good agreement with experimental data. The values of $`J^{(1)}`$ calculated with D1 and D1P are appreciable below both the experiment and the results obtained with D1S, see Fig. 10b. This is mainly due to smaller $`J^{(1)}`$ values for neutrons. At low rotational frequencies, the $`J^{(2)}`$ values obtained with D1 and D1P are lower than both experimental data and the ones obtained with D1S. However, the increase of $`J^{(2)}`$ as a function of rotational frequency is larger in the calculations with D1 and D1P and thus at $`\mathrm{\Omega }_x0.32`$ MeV they become larger than both experiment and the values calculated with D1S. In the neutron band crossing region at $`\mathrm{\Omega }_x0.42`$ MeV, the $`J^{(2)}`$ values calculated with D1 and D1P are larger than both the ones obtained with D1S and the experiment. It is interesting to note that all three sets give very similar neutron band crossing frequencies. Similar to the case of $`J^{(1)}`$, the differences between $`J^{(2)}`$’s calculated with D1 and D1P on the one side and with D1S on the other side are connected mainly with different alignment patterns in neutron subsystem. Calculated neutron and proton pairing energies $`E_{pairing}^{\nu ,\pi }`$ are displayed in Fig. 11b. The results of the calculations with different sets show a similar behavior as a function of rotational frequency, but the absolute values of pairing energies strongly depend on the parametrization of the Gogny force. The strongest pairing correlations are provided by the D1 set, while the smallest ones are obtained with D1S. The pairing energies calculated with D1P are in between the ones obtained with D1S and D1. Comparing transition quadrupole moments $`Q_t`$ obtained with different sets (see Fig. 11a), one can see that (i) the D1 and D1P sets provide very similar values of $`Q_t`$ and (ii) at high rotational frequencies the values of $`Q_t`$ calculated with all three sets come very close to each other reflecting the decrease of pairing correlations. An alternative way to study the role of pairing correlations for different physical observables is to look how the results of the calculations are affected when the strength of the Gogny force is changed. This is done by introducing the scaling factor $`f`$ into the Gogny force (see Eq. (30)). Such an investigation is in part motivated by the fact that in many studies performed within the relativistic Hartree-Bogoliubov theory with Gogny forces in the pairing channel and with no particle number projection, the strength of the Gogny force is increased by the factor 1.15, see for example Ref. . The results of the CRHB calculations for the lowest SD configuration in <sup>192</sup>Hg with the scaling factors $`f=0.9`$, $`f=1.0`$ and $`f=1.1`$ for the D1S Gogny force are shown in Figs. 12 and 13. The effect of the scaling of the Gogny force is especially drastic on the pairing and rotational properties of the configuration under investigation (see Figs. 13b and 12). The increase (decrease) of the strength of the Gogny force by 10% leads to approximately twofold increase (decrease) of the absolute values of proton and neutron pairing energies, see Fig. 13b. On the contrary, the increase (decrease) of the strength of the Gogny force leads to a significant decrease (increase) of the kinematic moment of inertia (Fig. 12b). One should note that the impact of the change of the strength of the Gogny force on the proton and neutron $`J^{(1)}`$ values is different. The effect of the scaling of the Gogny force on the dynamic moment of inertia is more complicated. While at low rotational frequencies it is similar to the one seen for the kinematic moment of inertia, it is completely different in the paired band crossing regions as seen in the proton, neutron and the total dynamic moments of inertia (Fig. 12a). Contrary to the low frequency range, stronger pairing correlations lead to larger $`J^{(2)}`$ values. In addition, as it is seen on the example of the neutron subsystem, stronger pairing correlations (a larger strength of the Gogny force) lead to the shift of the neutron paired band crossing to higher frequencies and make it sharper. A larger strength of the Gogny force leads also to smaller values of $`Q_t`$ at low $`\mathrm{\Omega }_x`$ and to larger $`Q_t`$ at high $`\mathrm{\Omega }_x`$, see Fig. 13a. In addition, different strengths and different sets of the Gogny force lead to different quasiparticle spectra in the vicinity of the Fermi level. In part, this effect is caused by the different calculated equilibrium deformations. It is reasonable to expect that this dependence of the quasiparticle spectra upon the parametrization and the strength of the Gogny force will have a pronounced impact on the band crossing frequencies in the SD configurations in odd and odd-odd nuclei. ### C Concluding remarks. The difference between the transition quadrupole moments $`Q_t`$ calculated with different sets of the RMF forces using the same set of the Gogny force is appreciable (Figs. 9c,d). On the contrary, the difference between $`Q_t`$’s obtained with three different sets (Fig. 11) or with three different scalings of the strength (Fig. 13a) of the Gogny force is smaller reflecting the fact that the equilibrium deformations are mostly defined by the properties of effective interaction in the particle-hole channel. The experimental uncertainties on transition quadrupole moments (see discussion in Ref. ) prevent the use of calculated $`Q_t`$’s for the selection of a better parametrization of the RMF Lagrangian for the particle-hole channel and the Gogny force for the particle-particle channel. Thus this selection can be based only on the rotational properties of the SD bands under study such as the kinematic and dynamic moments of inertia which are very accurately defined in experiment. Comparing different parametrizations of the Gogny force, it is clear that the D1 and D1P sets in connection with the NL1 force provide too strong pairing correlations and give very similar results for kinematic and dynamic moments of inertia, which deviate appreciable from experiment. The results of the calculations with the NL3 and NLSH forces in conjuction with the D1S set of the Gogny force also lead to considerable deviations from experiment for $`J^{(1)}`$ and $`J^{(2)}`$. It is then reasonable to expect that the use of the D1 and D1P sets of the Gogny force in conjuction with NL3 and NLSH will lead to even larger deviations from experiment for kinematic and dynamic moments of inertia. Thus one can conclude (see also Ref. ) that only the NL1 force in conjuction with the D1S set of the Gogny force and APNP(LN) will lead to a reasonable description of rotational properties of SD bands in the $`A190`$ mass region. ## V Properties of yrast SD bands in even-even nuclei It is our believe that the strong and the weak points of a theoretical approach can be determined only by a systematic comparison between experiment and theory. In the present Section the results of a systematic investigation of yrast SD bands in even-even nuclei of the $`A190`$ mass region will be presented. This investigation covers all even-even nuclei in this region in which SD bands have been observed so far, namely, <sup>190,192,194</sup>Hg, <sup>192,194,196,198</sup>Pb and <sup>198</sup>Po. First, the results for Hg and Pb isotopes with $`N=110,112`$ and 114 will be presented in detail. Partial results for these nuclei have been already presented in Ref. . Then the results for <sup>198</sup>Pb and <sup>198</sup>Po will be discussed for each nucleus separately. All the calculations presented in this Section have been performed with the force NL1 for the RMF Lagrangian and the set D1S for the Gogny force in the $`pp`$-channel and approximate particle number projection by means of the Lipkin-Nogami method (APNP(LN)) has been used. ### A The $`N=110,112`$ and $`114`$ Hg and Pb nuclei The calculated total, neutron and proton dynamic and kinematic moments of inertia are shown in Figs. 14, 15 and compared with experiment. One can see that a very successful description of dynamic moments of inertia of experimental bands is obtained in the calculations without adjustable parameters (Fig. 14). When comparing calculated and experimental kinematic moments of inertia, one should keep in mind that only yrast SD bands in <sup>194</sup>Pb and <sup>194</sup>Hg are definitely linked to the low-spin level scheme . In addition, there is a tentative linking of the SD band in <sup>192</sup>Pb . On the contrary, at present the yrast SD bands in <sup>190,192</sup>Hg and <sup>196</sup>Pb are not linked to the low-spin level scheme yet. Thus some spin values consistent with the signature of the calculated lowest SD configuration should be assumed for the experimental bands when a comparison is made with respect of the kinematic moment of inertia $`J^{(1)}`$. Taking into account that kinematic moments of inertia of linked SD bands in <sup>194</sup>Pb and <sup>194</sup>Hg and tentatively linked SD band in <sup>192</sup>Pb are well described in the calculations, the spin values of unlinked bands can be obtained by comparing the calculated values of $`J^{(1)}`$ with experimental ones under different spin assignments. Such a comparison (see Ref. for detailed figures) leads to the spin values $`I_0`$ for the lowest states in SD bands listed in Table III. Under these spin assignments, the CRHB+LN calculations rather well describe ’experimental’ kinematic moments of inertia of SD bands in <sup>192,196</sup>Pb and <sup>194</sup>Hg (Fig. 15). Alternative spin assignments, which are different from the ones given in Table III by $`\pm 2\mathrm{}`$, can be ruled out since they (open circles in Fig. 15) lead to considerable deviations from the results of the calculations. The increase of kinematic and dynamic moments of inertia in this mass region can be understood in the framework of the CRHB+LN theory as emerging predominantly from a combination of three effects: the gradual alignment of a pair of $`j_{15/2}`$ neutrons, the alignment of a pair of $`i_{13/2}`$ protons at a somewhat higher frequency, and decreasing pairing correlations with increasing rotational frequency, see also Sect. III. The interplay of alignments of neutron and proton pairs is more clearly seen in Pb isotopes where the calculated $`J^{(2)}`$ values show either a small peak (for example, at $`\mathrm{\Omega }_x0.44`$ MeV in <sup>192</sup>Pb, see Fig. 14) or a plateau (at $`\mathrm{\Omega }_x0.4`$ MeV in <sup>196</sup>Pb, see Fig. 14). With increasing rotational frequency, the $`J^{(2)}`$ values determined by the alignment in the neutron subsystem decrease. The maximum in $`J^{(2)}`$ of the neutron subsystem is reached at $`\mathrm{\Omega }_x0.44`$, 0.41, 0.39, 0.425 and 0.39 MeV in <sup>192,194,196</sup>Pb and <sup>194,196</sup>Hg, respectively. The decrease in these frequencies with increasing $`N`$ within each isotope chain correlates with the decrease of the transition quadrupole moments $`Q_t`$ (Fig. 4 in Ref. ). However, the fact that the maximum in neutron $`J^{(2)}`$ is obtained at the same frequencies in Hg and Pb isotones, which have $`Q_t`$ values differing by $`11.5`$ $`e`$b (Fig. 4 in Ref. ), indicates that the alignment process is very complicated and depends not only on the equilibrium deformation but also on the position of the Fermi level. The decrease of neutron $`J^{(2)}`$ at frequencies higher than $`\mathrm{\Omega }_x0.4`$ MeV is in part compensated by the increase of proton $`J^{(2)}`$ due to the alignment of the $`i_{13/2}`$ proton pair. This leads to the increase of the total $`J^{(2)}`$-value at $`\mathrm{\Omega }_x0.45`$ MeV in the Pb isotopes, while no such increase has been found in the calculations after the peak up to $`\mathrm{\Omega }_x=0.5`$ MeV in <sup>192</sup>Hg, see Fig. 14. Thus one can conclude that the shape of the peak (plateau) in total $`J^{(2)}`$ in the band crossing region is determined by a delicate balance between alignments in the proton and neutron subsystems which depends on deformation, rotational frequency and Fermi energy. It is also of interest to mention that the sharp increase in $`J^{(2)}`$ of the yrast SD band in <sup>190</sup>Hg is also reproduced in the present calculations. In the calculations, this increase is due to a two-quasiparticle alignment associated with the $`\nu [761]3/2`$ orbital. One should note that the calculations slightly overestimate the magnitude of $`J^{(2)}`$ at the highest observed frequencies. Possible reasons could be the deficiencies either of the Lipkin-Nogami method or of the cranking model in the band crossing region or both of them. In addition, the calculations do not reproduce the sudden decrease in $`J^{(2)}`$ at the bottom of the SD band in <sup>192</sup>Pb, the origin of which is not understood so far, see Ref. . In order to compare relative properties of the calculated and experimental dynamic moments of inertia in more detail we tilt them by extracting the frequency dependent term from $`J^{(2)}`$. The resulting quantities ($`J^{(2)}130\mathrm{\Omega }_x`$) are shown in Fig. 16 for Hg (top panel) and Pb (middle panel) isotopes as well as for $`N=112`$ (bottom panel) isotones. Since the difference of the dynamic moments of inertia of two bands is proportional to the derivative of the effective (relative) alignment $`i_{eff}`$ of these bands (see Refs. ) for details), the effective alignments of compared bands will also be presented here. The relative properties of the dynamic moments of inertia of the yrast SD bands in <sup>192,194</sup>Hg nuclei are very well reproduced in the calculations. Both in calculations and in experiment, $`J^{(2)}`$(<sup>192</sup>Hg)$``$$`J^{(2)}`$(<sup>194</sup>Hg) at $`\mathrm{\Omega }_x0.3`$ MeV, while at higher frequencies $`J^{(2)}`$(<sup>194</sup>Hg)$``$$`J^{(2)}`$(<sup>192</sup>Hg) (Figs. 16a,b). This result correlates with the fact that effective alignment $`i_{eff}`$ in the <sup>192</sup>Hg/<sup>194</sup>Hg pair and especially its slope is also reproduced rather well in the calculations (see Fig. 17d). On the contrary, while the properties of $`J^{(2)}`$(<sup>190</sup>Hg) relative to $`J^{(2)}`$(<sup>192</sup>Hg) are reasonably well reproduced at frequencies $`\mathrm{\Omega }_x0.3`$ MeV, the situation is different at lower frequencies. There the difference between $`J^{(2)}`$ values stays almost constant around 6 MeV<sup>-1</sup> in experiment, while it is decreasing to zero at $`\mathrm{\Omega }_x0.2`$ MeV in the calculations (see Figs. 16a,b). This feature correlates with the fact that the effective alignment in this pair of SD bands is not well reproduced in the calculations (see Fig. 17b). The dynamic moments of inertia of yrast SD bands in <sup>192,194</sup>Pb are almost identical in experiment. This feature (Figs. 16c,d) and the effective alignment in the <sup>192</sup>Pb/<sup>194</sup>Pb pair (Figs. 17a) are rather well reproduced in the calculations. In experiment, the dynamic moments of inertia of SD bands in <sup>194,196</sup>Pb are almost identical at $`\mathrm{\Omega }_x0.1`$ MeV. With increasing $`\mathrm{\Omega }_x`$, $`J^{(2)}`$(<sup>196</sup>Pb) decreases below $`J^{(2)}`$(<sup>194</sup>Pb) with a maximum difference between them of $`5`$ MeV<sup>-1</sup> being reached at $`\mathrm{\Omega }_x0.22`$ MeV and then this difference becomes smaller up to $`\mathrm{\Omega }_x0.32`$ MeV where the dynamic moments of inertia of both bands coincide (Fig. 16c). At even higher frequencies, $`J^{(2)}`$(<sup>196</sup>Pb)$``$$`J^{(2)}`$(<sup>194</sup>Pb). The calculations reasonably well reproduce the general features, however, somewhat underestimate the difference between $`J^{(2)}`$’s of these two bands at medium frequencies and overestimate this difference at highest observed frequencies. The experimental effective alignment in the <sup>194</sup>Pb/<sup>196</sup>Pb pair is very well reproduced at low frequencies, while the difference between experiment and calculations is somewhat larger at high frequencies (Fig. 17c). The experimental dynamic moments of inertia of SD bands in <sup>194</sup>Pb and <sup>192</sup>Hg are identical at $`\mathrm{\Omega }_x0.2`$ MeV, while at higher frequencies the condition $`J^{(2)}`$(<sup>194</sup>Pb)$``$$`J^{(2)}`$(<sup>192</sup>Hg) holds (Fig. 16e). The relative properties of dynamic moments of inertia of these two bands at $`\mathrm{\Omega }_x0.2`$ MeV are rather well reproduced in the calculations (Fig. 16f). At lower frequencies, there is however some difference between the calculated $`J^{(2)}`$ values for these two bands of $`2`$ MeV<sup>-1</sup>. The effective alignment in the <sup>192</sup>Hg/<sup>194</sup>Pb pair and especially its slope as a function of $`\mathrm{\Omega }_x`$ is well reproduced in the calculations (Fig. 17e). The effective alignments between other pairs of SD bands which have not been discussed before are also shown in Fig. 17g,f,i. It is seen that the effective alignment in the <sup>194</sup>Hg/<sup>194</sup>Pb pair is reasonably well reproduced in the calculations, although they somewhat overestimate the absolute value of $`i_{eff}`$. Since the spins of these two bands are determined experimentally, this result can be considered as a first direct justification of the reliability of the effective alignment approach used frequently for the configuration and spin assignments of SD bands in different mass regions, see Refs. for details. Comparing the experimental and effective alignments (Fig. 17) and taking into account that compared bands differ by at least 2 particles, one can conclude that considerable deviations from experiment are seen only in the cases of the effective alignments in the <sup>190</sup>Hg/<sup>192</sup>Pb, <sup>194</sup>Hg/<sup>196</sup>Pb and <sup>190</sup>Hg/<sup>192</sup>Hg pairs (Figs. 17g,f,b). In order to find are observed deviations from experiment related to the particle-hole or the particle-particle channel of the CRHB theory the investigation of SD bands in odd nuclei of this mass region are needed. Such an investigation is in progress and its results will be reported later. The investigation of relative alignments of SD bands in this mass region has been performed in non-relativistic approaches such as the total routhian surface Strutinsky-type approach based on the Woods-Saxon potential and Skyrme-Hartree-Fock-Bogoliubov approach in Ref. . Similar to our case, some discrepancies between theory and experiment have been found. The calculated neutron and proton pairing energies $`E_{pairing}^{\nu ,\pi }`$ are summarized in Fig. 18. In all nuclei, they decrease with increasing rotational frequency reflecting the quenching of pairing correlations due to the Coriolis antipairing effect. In addition, neutron pairing is stronger than proton pairing. Neutron pairing energies in <sup>190,192</sup>Hg almost coincide up to the band crossing seen in <sup>190</sup>Hg. A similar situation exists also in <sup>192,194</sup>Pb where the difference between calculated neutron pairing energies does not exceed 0.2 MeV. These energies are smaller than the ones in <sup>190,192</sup>Hg by $`0.5`$ MeV at $`\mathrm{\Omega }_x=0.0`$ MeV, while at high rotational frequencies neutron pairing energies are similar in these Pb and Hg nuclei. On the contrary, neutron pairing energies in the $`N=114`$ Hg and Pb nuclei are larger than the ones in the nuclei with $`N=112`$ by 0.5-0.8 MeV dependent on the rotational frequency and they do coincide at $`\mathrm{\Omega }_x0.3`$ MeV. A similar trend is seen also for proton pairing energies where, however, the difference between the pairing energies in different nuclei is smaller than in the case of neutron pairing energies and it does not exceed 0.5 MeV. ### B The nucleus <sup>198</sup>Pb The results of the calculations for the lowest SD band in <sup>198</sup>Pb are shown in Figs. 19 and 20. The calculated kinematic moment of inertia agrees reasonably well with experiment up to $`\mathrm{\Omega }_x0.32`$ MeV, while considerable disagreement is seen at higher frequencies. The increase of calculated $`J^{(1)}`$ as a function of $`\mathrm{\Omega }_x`$ is larger in <sup>198</sup>Pb than in <sup>196</sup>Pb (Fig. 19b). As a result, the calculated dynamic moment of inertia in <sup>198</sup>Pb is larger than the one in <sup>196</sup>Pb at $`\mathrm{\Omega }_x0.2`$ MeV and smaller at $`\mathrm{\Omega }_x0.2`$ MeV. On the contrary, the experimental data show the opposite trend for the dynamic moment of inertia of the <sup>198</sup>Pb band having a much smaller increase as a function of rotational frequency (Fig. 19a). Thus the results of the calculations for <sup>198</sup>Pb do not reproduce neither absolute rotational properties of SD band nor their relative properties with respect to SD band in <sup>196</sup>Pb. A similar problem with the reproduction of the properties of the <sup>198</sup>Pb band exists also in the cranked Nilsson-Strutinsky Lipkin-Nogami calculations presented in Refs. . The investigation of neighbouring odd nuclei within the CRHB+LN theory is needed in order to understand better the origin of these problems. The calculated values of $`Q_t`$ for SD band in <sup>198</sup>Pb are by $`1`$ $`e`$b smaller than in <sup>196</sup>Pb (see Fig. 20a) which is in agreement with the decrease of $`Q_t`$ with increasing $`N`$ seen in the lighter Pb isotopes (see Fig. 4 in Ref. ). Precise measurements of the relative transition quadrupole moments of the yrast SD bands in <sup>196,198</sup>Pb nuclei and a theoretical study of neighboring odd nuclei can be useful for the understanding of the present problems with the description of $`J^{(1)}`$ and $`J^{(2)}`$. Pairing energies in <sup>198</sup>Pb are somewhat larger than in <sup>196</sup>Pb (Fig. 20b) which agrees with the trend of the increase of pairing correlations within the isotopic chain with increasing neutron number $`N`$ (see Section V A). ### C The nucleus <sup>198</sup>Po The CRHB+LN calculations for the lowest SD configuration in <sup>198</sup>Po very well describe experimental dynamic and kinematic moments of inertia of the yrast SD band in this nucleus (Fig. 21a). At $`\mathrm{\Omega }_x0.34`$ MeV, the calculations predict sharp increase in dynamic moment of inertia caused mainly by the alignment of the lowest proton hyperintruder $`\pi [770]1/2`$ orbital (see Fig. 22). The absolute values of proton and neutron pairing energies and their behavior as a function of rotational frequency (Fig. 21b) are similar to the ones seen in other nuclei. A specific feature of this nucleus, which has not been observed in other nuclei, is considerable increase of the transition quadrupole moment $`Q_t`$ (by $`3`$ $`e`$b) with increasing rotational frequency (Fig. 21c). At $`\mathrm{\Omega }_x=0.0`$ MeV, the $`Q_t`$ values in <sup>198</sup>Po are larger than the ones in isotonic <sup>196</sup>Pb by $`1.0`$ $`e`$b (see Fig. 21c and Fig. 4 in Ref. ). One should note that the results of the GCM+GOA calculations based on the Gogny force also show the same feature . Finally, the experimental effective alignment in the <sup>196</sup>Pb/<sup>198</sup>Po pair is reasonably well described in the calculations (Fig. 21d). ### D Mass hexadecupole moments $`Q_t`$ Experimental information on mass hexadecupole moments $`Q_{40}`$ is not available so far for SD bands in any mass region. Thus only the results of the calculations for this quantity are presented in Fig. 23. These results have been obtained with the NL1 force for the RMF Lagrangian and the set D1S for the Gogny force if it is not specified otherwise. Note that we do not show the results of the calculations obtained with either different parametrizations or different scalings of the Gogny force. The $`Q_{40}`$ values calculated with no pairing (dotted lines in Figs. 23a,b) show a gradual decrease with an increase of rotational frequency. A similar trend is also seen in the results of the calculations with no APNP(LN) (solid lines with solid squares in Figs. 23a,b). The sharp change of the slope of the $`Q_{40}(\mathrm{\Omega }_x)`$ curve seen in <sup>194</sup>Pb at $`\mathrm{\Omega }_x0.35`$ MeV (Fig. 23b) is related to a sharp band crossing associated with the collapse of proton pairing correlations (see Sect. III). The results of the calculations with APNP(LN) (open symbols in Fig. 23) show a somewhat different trend. With the exception of <sup>198</sup>Po, the calculated $`Q_{40}`$ values stay nearly constant or smoothly increase with increasing rotational frequency up to $`\mathrm{\Omega }_x0.35`$ MeV. Above this frequency they decrease with increasing $`\mathrm{\Omega }_x`$. In <sup>198</sup>Po, the $`Q_{40}`$ values increase with increasing $`\mathrm{\Omega }_x`$ in the whole calculated frequency range (Fig. 23c). This increase is especially pronounced in the band crossing region. Within the isotopic chain the increase of neutron number $`N`$ leads to the decrease of $`Q_{40}`$. The $`Q_{40}`$ values increase within isotonic chain with increasing proton number $`Z`$. The results of the calculations with NL3 and NLSH lead to smaller values of $`Q_{40}`$ compared with the ones obtained with NL1 force (Fig. 23a,b). The features discussed above are very similar to the ones obtained for the transition quadrupole moment $`Q_t`$, see the discussion in previous sections. ### E Particle number fluctuation $`(\widehat{\mathrm{\Delta }N})^2`$ The basis assumption behind the Kamlah expansion to second order used in the derivation of the Lipkin-Nogami method is that the system is well pair-correlated which means that the particle number fluctuation in the unprojected wave function $`(\mathrm{\Delta }\widehat{N})^2`$ is large. These quantities obtained in the CRHB+LN calculations with the NL1 and NL3 forces for the RMF Lagrangian and the D1S set for the Gogny force are shown in Fig. 24 for all nuclei studied in the present manuscript. The particle number fluctuations decrease more or less smoothly with increasing rotational frequency indicating the quenching of pairing correlations due to the Coriolis antipairing effect. One should note that even at the highest rotational frequencies these fluctuations remain reasonably large thus indicating that the approximate particle number projection by means of the Lipkin-Nogami method still remains within the applicability range of the Kamlah expansion. The values of $`(\mathrm{\Delta }\widehat{N})^2`$ in neutron and proton subsystems correlate with the pairing energies $`E_{pairing}`$ calculated in these subsystems. For example, in the calculations with the NL1 force the particle number fluctuations are larger for neutrons than for protons (see Fig. 24) which correlates with the fact that the absolute values of pairing energies are larger for neutrons (see Figs. 18, 20b and 21b). The situation is somewhat different in the calculations with the NL3 force, where at medium and high rotational frequencies the proton subsystem is more pair-correlated than the neutron one as reflected in the particle number fluctuations (Fig. 24) and pairing energies (Fig. 9a,b). ## VI Conclusions The formalism of the Cranked Relativistic Hartree-Bogoliubov theory with and without approximate particle number projection before variation by means of the Lipkin-Nogami method is presented in detail. The relativistic mean field theory is used in the particle-hole channel of this theory, while a non-relativistic finite range two-body force of Gogny type is employed in the particle-particle (pairing) channel. Considering that the pairing is a genuine non-relativistic effect which plays a role only in the vicinity of the Fermi surface, the use of the best non-relativistic force in the pairing channel seems well justified. Its applicability to the description of rotating nuclei and the main features of this theory have been studied on the example of the yrast superdeformed bands observed in even-even nuclei of the $`A190`$ mass region. The main conclusions emerging from this study are the following: (i) The calculations without particle number projection do provide only a poor description of experimental rotational features such as the kinematic $`J^{(1)}`$ and the dynamic $`J^{(2)}`$ moments of inertia. The calculated kinematic moments of inertia are larger than the experimental values. The same is also true for the dynamic moments of inertia at low and medium rotational frequencies. The calculations without particle number projection lead to a unphysical collapse of pairing correlations as has been seen in the proton subsystem of <sup>192</sup>Hg and <sup>194</sup>Pb. As was shown by subsequent calculations with particle number projection these problems are related to the poor treatment of pairing correlations. (ii) Approximate particle number projection by means of the Lipkin-Nogami method considerably improves an agreement with experiment. The correlations induced by the Lipkin-Nogami method produce the desired effects, namely, (a) to increase the strength of the pairing correlations, (b) to diminish the kinematic moments of inertia at all frequencies, (c) to diminish the dynamic moments of inertia at low and medium rotational frequencies and (d) to delay proton and neutron alignments to higher frequencies. In addition, there is no collapse of pairing correlations in the whole rotational frequency range under investigation. Systematic calculations with the NL1 force for the RMF Lagrangian and the D1S set for the Gogny force have been performed for all even-even nuclei in which SD bands have been observed so far. With an exception of <sup>198</sup>Pb, an excellent description of the rotational properties of yrast SD bands such as the dynamic and kinematic moments of inertia is obtained in a way free from adjustable parameters. It was concluded that the investigation of the structure of SD bands in neighboring odd nuclei is needed in order to better understand the problems seen in this nucleus. (iii) It has been investigated how much the results of calculations with approximate particle number projection by means of the Lipkin-Nogami method depend on the parametrization of the RMF Lagrangian and the Gogny force. It was found that the combination of the NL1 set for the RMF Lagrangian and the D1S set for the Gogny force produce very good agreement with experimental rotational properties. The D1 and D1P sets of the parameters for the Gogny force produce too strong pairing correlations and thus fail to describe properly the rotational properties of SD bands. An unexpected result of the present investigation is the fact that the NL3 force, which is believed to be the best RMF force for the description of nuclear properties far from beta-stability region, provides a less accurate description of the rotational properties of SD bands in the $`A190`$ mass region compared with the force NL1. This is most likely related to the single-particle spectra produced by this force in the SD minimum. It is possible that the improvement of the description of the isospin properties far from beta-stability region obtained in the NL3 force as compared with the NL1 force is reached at the cost of some worsening of the description of the single-particle spectra close to beta-stability region. (iv) The dependence of the results on the strength of the Gogny force in the pairing channel has been studied in the calculations with approximate particle number projection by means of the Lipkin-Nogami method. It was found that a change of the strength by $`\pm 10`$% has significant impact on the rotational properties of SD bands such as the kinematic and dynamic moment of inertia. Contrary to previous investigations within the Relativistic Hartree-Bogoliubov theory at no rotation and with no particle number projection, where the strength of the Gogny force has been increased by factor 1.15, here a very good description of rotational properties has been obtained with no modification of the strength of the Gogny force. (v) The results of the calculations with particle number projection indicate the general trend of a decrease of the average transition quadrupole ($`Q_t`$) and mass hexadecupole $`(Q_{40}`$) moments in the isotopic chain with increasing neutron number $`N`$ and an increase of these quantities within the isotonic chain with increasing proton number $`Z`$. (vi) Neutron and proton pairing energies in all the calculated nuclei decrease with increasing rotational frequency reflecting the quenching of pairing correlations due to the Coriolis anti-pairing effect. The pairing energies $`E_{pairing}`$ calculated in the NL1+D1S+LN scheme do not show considerable variations as a function of $`Z`$ and $`N`$: the maximum difference between pairing energies calculated in two different nuclei is around 16%. A smooth increase of absolute values of pairing energies with increasing $`N`$ is observed in the isotopic Pb and Hg chains which correlates with the decrease of the transition quadrupole moments $`Q_t`$. The size of the pairing energies strongly depends on the parametrization of the Gogny force and the RMF Lagrangian, as well as on the strength of the Gogny force. (vii) The difference between the experimental dynamic moments of inertia of two bands and the effective alignment $`i_{eff}`$ between these bands is reasonably well reproduced in most of the cases. Further investigation of neighboring odd nuclei is needed in order to find the origin of the remaining discrepancies. Finally, the present work should be considered as one of the first steps in the investigation of rotating nuclei in the pairing regime within the framework of the RMF theory. Different tasks definitely lie ahead. For example, an investigation of the rotational bands based on one- and multi-quasiparticle configurations in the $`A190`$ mass region is mandatory in order to see how the present theory can reproduce the effects connected with the blocking of one or several quasiparticle orbitals. One can expect that the Lipkin-Nogami method is a reasonably good approximation to the exact particle number projection in the regimes of strong pairing correlations as it holds in the case of the $`A190`$ mass region. It remains to investigate if this method is also a good approximation in the regimes of weak pairing correlations typical at high rotational frequencies in SD bands of other regions of the periodic table, such as the $`A60`$, $`A130`$ and $`A150`$ mass regions. ## VII Acknowledgments A.V.A. acknowledges support from the Alexander von Humboldt Foundation. This work is also supported in part by the Bundesministerium für Bildung und Forschung under the project 06 TM 875.
warning/0001/math0001060.html
ar5iv
text
# Special Lagrangian Geometry in irreducible symplectic 4-folds ## 1 Introduction Under the flourishing research activity on D-branes in string theory, the role of special Lagrangian submanifolds in physics has become more and more relevant (see for example ) untill it was eventually conjectured in that they can be considered as the cornerstones of the mirror phenomenon. Indeed, D-branes are special Lagrangian submanifolds equipped with a flat $`U(1)`$ line bundle. In physical literature, special Lagrangian submanifolds of the compactification space are related to physical states which retain part of the vacuum supersymmetry: for this reason they are often called supersymmetric cycles or BPS states. Despite their importance, there are very few explicit examples of special Lagrangian submanifolds, especially in Calabi-Yau 3-folds. However, in an irreducible symplectic 4-fold (realized as a hyperkaehler manifold) we have a complete control of the special Lagrangian geometry of its submanifolds, via a sort of hyperkaehler trick ; moreover this enables us to prove that special Lagrangian submanifolds retain part of the rigidity of complex submanifolds. We first recall the following: Definition 1.1: A complex manifold $`X`$ is called irreducible symplectic if it satisfies the following three conditions: 1) $`X`$ is compact and Kaehler; 2) $`X`$ is simply connected; 3) $`H^0(X,\mathrm{\Omega }_X^2)`$ is spanned by an everywhere non-degenerate 2-form $`\omega `$. In particular, irreducible symplectic manifolds are special cases of Calabi-Yau manifolds (the top holomorphic form which trivializes the canonical line bundle is given by a suitable power of the holomorphic 2-form $`\omega `$). In dimension 2, K3 surfaces are the only irreducible symplectic manifolds, and indeed irreducible symplectic manifolds appear as higher-dimensional analogues of K3 surfaces, as strongly suggested in . Unfortunately, up to now there are very few explicit examples of irreducible symplectic manifolds. Indeed almost all known examples turn out to be birational to two standard series of examples: Hilbert schemes of points on K3 surfaces and generalized Kummer varieties (both series were first studied in ), but quite recently O’Grady has constructed irreducible symplectic manifolds which are not birational to any of the elements of the two groups (see ). Finally, let us recall from the following: Definition 1.2: Let $`X`$ be a Calabi-Yau n-fold, with Kaehler form $`\omega `$ and holomorphic nowhere vanishing n-form $`\mathrm{\Omega }`$. A (real) n-dimensional submanifold $`j:\mathrm{\Lambda }X`$ of $`X`$ is called special Lagrangian if the following two conditions are satisfied: 1) $`\mathrm{\Lambda }`$ is Lagrangian with respect to $`\omega `$, i.e. $`j^{}\omega =0`$; 2) there exists a multiple $`\mathrm{\Omega }^{}`$ of $`\mathrm{\Omega }`$ such that $`j^{}\mathrm{Im}(\mathrm{\Omega }^{})`$=0; one can prove (see ) that both conditions are equivalent to: 1’) $`j^{}\mathrm{Re}(\mathrm{\Omega }^{})=Vol_g(\mathrm{\Lambda })`$. The condition $`1^{})`$ in the previous definition means that the real part of $`\mathrm{\Omega }^{}`$ restricts to the volume form of $`\mathrm{\Lambda }`$, induced by the Calabi-Yau Riemannian metric $`g`$. In this way special Lagrangian submanifolds are considered as a type of calibrated submanifolds (see for further details on this point). ## 2 Characterization of special Lagrangian submanifolds In this section we will describe all special Lagrangian submanifolds of an irreducible symplectic 4-fold $`X`$ (having fixed a Kaehler class $`[\omega ]`$ in the Kaehler cone). The key result is the following: Theorem 2.1: Every connected special Lagrangian submanifold of an irreducible symplectic 4-fold is also bi-Lagrangian, in the sense that it is Lagrangian with respect to two different symplectic structures. Proof: Let us a fix a Kaehler class on the irreducible symplectic 4-fold $`X`$. By Yau’s Theorem this determines a unique hyperkaehler metric $`g`$. Choose a hyperkaehler structure $`(I,J,K)`$ compatible with the metric $`g`$ (notice that the triple $`(I,J,K)`$ is not uniquely determined) and consider the associated symplectic structures $`\omega _I(.,.):=g(I.,.)`$, $`\omega _J(.,.):=g(J.,.)`$ and $`\omega _K(.,.):=g(K.,.)`$. Consider a special Lagrangian submanifold $`\mathrm{\Lambda }`$ in the complex structure $`K`$ (this is not restrictive, since $`(I,J,K)`$ is not uniquely determined); that is assume that $`\mathrm{\Lambda }`$ is calibrated by the real part of the holomorphic (in the structure $`K`$) 4-form: $$\mathrm{\Omega }_K:=\frac{1}{2!}(\omega _I+i\omega _J)^2.$$ (1) Notice that the real and immaginary part of $`\mathrm{\Omega }_K`$ are then given by: $$\mathrm{Re}(\mathrm{\Omega }_K)=\frac{1}{2}(\omega _I^2\omega _J^2)\mathrm{Im}(\mathrm{\Omega }_K)=\omega _I\omega _J.$$ (2) Obviously, by the property of being special Lagrangian we have that $`\mathrm{\Lambda }`$ is Lagrangian with respect to $`\omega _K`$. We will prove that having fixed the calibration, if $`\mathrm{\Lambda }`$ is not Lagrangian also with respect to $`\omega _I`$, then it is necessarily Lagrangian with respect to $`\omega _J`$. First we work locally and consider $`V:=T_p\mathrm{\Lambda }`$ ($`p\mathrm{\Lambda }`$), spanned by $`(w_1,w_2,w_3,w_4)`$. Since $`\mathrm{\Lambda }`$ is assumed not to be Lagrangian with respect to $`\omega _I`$, we have to deal with two cases. First case: $`V`$ is a symplectic vector space for the structure $`\omega _I`$. In this case we can choose a symplectic basis for $`V`$ and this can always be chosen to be of the form $`v_1,Iv_1,v_2,Iv_2`$. Then $`V`$ is Lagrangian in the symplectic structure $`\omega _J`$; indeed $`\omega _J(v_1,Iv_1)=g(Jv_1,Iv_1)=g(IJv_1,v_1)=\omega _K(v_1,v_1)=0`$; analogously for $`\omega _J(v_2,Iv_2)`$; $`\omega _J(v_1,Iv_2)=g(Jv_1,Iv_2)=\omega _K(v_1,v_2)=0`$ since $`v_1,v_2`$ belong to a Lagrangian subspace of $`\omega _K`$, and analogously for $`\omega _J(v_2,Iv_1)=\omega _K(v_2,v_1)=0`$. Thus $`V`$ is also Lagrangian for the symplectic structure $`\omega _J`$. Second case: $`V`$ is neither symplectic nor Lagrangian for the structure $`\omega _I`$. Notice $`V`$ can not be symplectic with respect to $`\omega _J`$, otherwise by the first case it would be Lagrangian in the strucutre $`\omega _I`$; moreover we can assume that $`V`$ is not Lagrangian with respect to $`\omega _J`$, otherwise there is nothing to prove. So in this case $`V`$ is neither Lagrangian nor symplectic in the structure $`\omega _I`$ and in the structure $`\omega _J`$. This means that $`V`$ contains a symplectic 2-plane $`\pi `$ with respect to $`\omega _I`$ and a symplectic 2-plane $`\rho `$ with respect to $`\omega _J`$. Indeed, consider $`v_1V`$; since $`V`$ is not Lagrangian in the structure $`\omega _I`$, there exists $`v_2V`$ such that $`\omega _I(v_1,v_2)0`$ and this implies that the vector subspace $`\pi `$ spanned by $`(v_1,v_2)`$ is a symplectic vector space for $`\omega _I`$, which can not be extended to all $`V`$. The same reasoning applies in the structure $`\omega _J`$. We prove that this can not happen, since it violates the calibration condition. We have to distinguish three different subcases according to the intersection of $`\pi `$ with $`\rho `$. First subcase: $`\pi `$ and $`\rho `$ have zero intersection. If this happens we can always choose a basis of $`V`$ of the form $`(v_1,Iv_1,v_2,Jv_2)`$. Write $`\pi `$ for the 2-plane spanned by $`v_1,Iv_1`$ and $`\rho `$ for that spanned by $`v_2,Jv_2`$, so that $`V=\pi \rho `$. Indeed, since $`V`$ is not Lagrangian with respect to $`\omega _I`$, it has to contain a symplectic 2-plane like $`\pi `$, and similarly for $`\rho `$ and $`\omega _J`$. Moreover, since $`V`$ is not symplectic with respect to $`\omega _I`$, it turns out that the symplectic 2-plane $`\pi `$ can not be completed to a symplectic basis of $`V`$, so that $`V`$ has to contain an isotropic 2-plane for $`\omega _I`$, which is $`\rho `$. The same reasoning (with the roles reversed) applies obviously to the symplectic structure $`\omega _J`$. Hence, in this case we have: $$2\mathrm{R}\mathrm{e}(\mathrm{\Omega }_K)|_V=(\omega _I^2\omega _J^2)(v_1,Iv_1,v_2,Jv_2)=\omega _I(v_1,Iv_1)\omega _I(v_2,Jv_2)$$ $$\omega _I(v_1,v_2)\omega _I(Iv_1,Jv_2)+\omega _I(v_1,Jv_2)\omega _I(Iv_1,v_2)\omega _J(v_1,Iv_1)\omega _J(v_2,Jv_2)+$$ $$\omega _J(v_1,v_2)\omega _J(Iv_1,Jv_2)\omega _J(v_1,Jv_2)\omega _J(Iv_2,v_2)=0,$$ using the defining relations of $`\omega _I,\omega _J,\omega _K`$, the quaternionic relation $`IJ=K`$, the invariance of $`g`$ and the fact that $`V`$ is Lagrangian with respect to $`\omega _K`$. So this subcase is not consistent with the calibration property. Second subcase: $`\pi `$ and $`\rho `$ have a 1-dimensional intersection spanned by a vector $`v_1`$. In this case we can choose a basis of $`V`$ of the form $`(v_1,Iv_1,Jv_1,w)`$ ($`\pi `$ is spanned by $`(v_1,Iv_1)`$, while $`\rho `$ is spanned by $`(v_1,Jv_1)`$). Again by the same computation of the previous subcase one shows that this configuration is not compatible with the calibration. Third subcase: Finally $`\pi `$=$`\rho `$ can not clearly happen, since otherwise one can choose a basis of $`\pi `$ equal to $`(v_1,Iv_1)`$, but then, in this basis $`\omega _J`$ is identically vanishing, contrary to the assumption that $`\rho =\pi `$ is a symplectic 2-plane also for $`\omega _J`$. Since the second case can never happen $`V`$ has to be Lagrangian also with respect to $`\omega _J`$. Up to now, we have worked only locally; to conclude the proof it is necessary to show that if $`T_p\mathrm{\Lambda }`$ is Lagrangian with respect to $`\omega _J`$, then it can not be possible that $`T_q\mathrm{\Lambda }`$ is Lagrangian with respect to $`\omega _I`$, for a different $`q\mathrm{\Lambda }`$. Notice that any tangent space to $`\mathrm{\Lambda }`$ can not be Lagrangian with respect to both $`\omega _I`$ and $`\omega _J`$, otherwise it would violates the calibration condition. Consider now the following smooth sections of $`^2T^{}\mathrm{\Lambda }`$: $$\begin{array}{cccc}\alpha _{I,J}:& \mathrm{\Lambda }& & ^2T^{}\mathrm{\Lambda }\\ & p& & \omega _{I,J}|T_p\mathrm{\Lambda }\end{array}$$ and the zero section $`s_0:\mathrm{\Lambda }^2T^{}\mathrm{\Lambda }`$. Obviously, $`s_0(\mathrm{\Lambda })`$ is closed in $`^2T^{}\mathrm{\Lambda }`$, and by the previous reasoning $`\mathrm{\Lambda }`$ can be decomposed as $`\mathrm{\Lambda }=\alpha _I^1(s_0(\mathrm{\Lambda }))\alpha _J^1(s_0(\mathrm{\Lambda }))`$, that is as the disjoint union of two proper closed subsets. But this is clearly impossible, since $`\mathrm{\Lambda }`$ is connected, and this implies that one of the two closed subset is empty, so $`\mathrm{\Lambda }`$ is bi-Lagrangian. The previous theorem is important in view of the following: Corollary 2.1: Every (connected, compact and without border) special Lagrangian submanifold $`\mathrm{\Lambda }`$ of a hyperkaehler 4-fold $`X`$ can be realized as a complex submanifold, via hyperkaehler rotation of the complex structure of $`X`$. Proof: Let $`\mathrm{\Lambda }`$ be a special Lagrangian submanifold of $`X`$ in the complex structure $`K`$. Then by definition $`\mathrm{Re}(\mathrm{\Omega }_K)_{|\mathrm{\Lambda }}=\mathrm{Vol}_g(\mathrm{\Lambda })`$, but by the previous theorem, since $`\omega _J|_\mathrm{\Lambda }=0`$ this means: $$\mathrm{Vol}_g(\mathrm{\Lambda })=\frac{1}{2}_\mathrm{\Lambda }\omega _I^2.$$ (3) By Wirtinger’s theorem, since $`\mathrm{\Lambda }`$ is assumed to be compact and without border, condition (3) is equivalent to say that $`\mathrm{\Lambda }`$ is a complex submanifold of $`X`$, in the complex structure $`I`$, that is performing a hyperkaehler rotation. Notice that in the complex structure $`I`$, $`\mathrm{\Lambda }`$ is still a Lagrangian submanifold with respect to $`\omega _K`$ and $`\omega _I`$, so it is Lagrangian with respect to the holomorphic (in the structure $`I`$) 2-form $`\mathrm{\Omega }_I:=\omega _J+i\omega _K`$. Collecting the results so far proved, we can show that special Lagrangian submanifolds of $`X`$ are particularly rigid : Proposition 2.1: Any (connected, compact and without border) special Lagrangian submanifold $`\mathrm{\Lambda }`$ of a hyperkaehler 4-fold $`X`$ is real analytic. Proof: Let $`\mathrm{\Lambda }`$ be a special Lagrangian submanifold of $`X`$, having fixed some complex structure on $`X`$, let us say $`K`$; then, by Corollary 2.1 there exists a new complex structure, let us say $`I`$, in which $`\mathrm{\Lambda }`$ is holomorphic, that is, it is locally given by: $$f_1(z_1,\mathrm{},z_4)=0\mathrm{and}f_2(z_1,\mathrm{},z_4)=0.$$ Now observe that coming back to the original complex structure $`K`$, we induce an analytic change of coordinates from the holomorphic coordinates $`z^i`$ ($`I\frac{}{z^i}=i\frac{}{z^i}`$) to new holomorphic coordinates $`w^i`$ ($`K\frac{}{w^i}=i\frac{}{w^i}`$) such that locally: $$z^i=c_1w^i+c_2\overline{w}^i\overline{z}^i=d_1w^i+d_2\overline{w}^i,$$ (4) for some complex constants $`c_j,d_j`$. Thus in the complex structure $`K`$ the special Lagrangian submanifold $`\mathrm{\Lambda }`$ is given by $`f_j(c_1w^i+c_2\overline{w}^i,d_1w^i+d_2\overline{w}^i)=0`$ which is again the zero locus of a set of functions analytic in $`w^i,\overline{w}^i`$. Quite naturally, the action of the hyperkaehler rotation can be extended also to the holomorphic functions defined on complex submanifolds $`S`$ of $`X`$; in particular we have an action of the hyperkaehler rotation on the structure sheaf $`𝒪_S`$ (here, as always, we identify $`𝒪_S`$ with its direct image $`j_{}𝒪_S`$, where $`j:SX`$ is the holomorphic embedding). We are thus led to give the following: Definition 2.2: Let $`\mathrm{\Lambda }`$ be a special Lagrangian submanifold of a hyperkaehler 4-fold $`X`$ (in the complex structure $`K`$). Then we define the special Lagrangian structure sheaf $`_\mathrm{\Lambda }`$ as the sheaf obtained by the action of the hyperkaehler rotation on the structure sheaf $`𝒪_\mathrm{\Lambda }`$ of $`\mathrm{\Lambda }`$, as a complex Lagrangian submanifold of $`X`$, (in the structure $`I`$). ## 3 Concluding remarks It is important to remark that all previous results are true also for special Lagrangian submanifolds of K3 surfaces, but their proof is completely trivial in that case. Another observation is related to singular Lagrangian submanifolds: indeed, by the previous results, it turns out that we can also give examples of special Lagrangian subvarieties, obtained via hyperkaehler rotation of Lagrangian complex subvarieties. On the other hand, contrary to the case of the corresponding submanifolds, we can not expect that all special Lagrangian subvarieties are obtained in this way, and consequently we can not expect that all special Lagrangian subvarieties are real analytic. Indeed, there are examples (compare ) of singular special Lagrangian submanifold in $`C^n`$ which are only smooth, but not real analytic. The discussion about singular Lagrangian submanifolds leads us to comment on the mirror symmetry construction suggested in . Indeed, according to the recipe of , any Calabi-Yau $`X`$, admitting a mirror $`\widehat{X}`$, has a peculiar fibre space structure: on a physical ground it is argued that $`X`$ can be realized as the total space of a fibration in special Lagrangian tori. Unfortunately, there are very few examples of such realization: in particular, as far as we know, there is only one (partial) example for Calabi-Yau 3-folds of the so called Borcea-Voisin type (see ). Instead, in the case of irreducible symplectic projective manifolds the situation is completely different. Indeed, a recent result of Matsushita (see and ) shows that for any fibre space structure $`f:XB`$ of a projective irreducible symplectic manifold $`X`$, with projective base $`B`$, the generic fibre $`f^1(b)`$ is an Abelian variety (up to finite unramified cover), and it is also Lagrangian with respect to the non degenerate holomorphic 2-form $`\mathrm{\Omega }`$; moreover, in the case of 4-folds one can prove that the generic fibre is an Abelian surface and $`f`$ is equidimensional, (i.e. all irreducible components of the fibres have the same dimension). By Corollary 2.1 it turns out that this fibre space structure can also be realized as a special Lagrangian torus fibration; moreover, in this case all special Lagrangian fibres, even the singular ones, are analytic, since they are obtained by performing a hyperkaehler rotation starting from Lagrangian Abelian surfaces. So, in these cases, we have special Lagrangian torus fibration in which all fibres are analytic: one can hope to understand the degeneration types of singular special Lagrangian tori, moving from these constructions. Explicit examples of projective irreducible symplectic 4-folds, fibered over a projective base have been constructed by Markuschevich in and . One of this constructions is the following: consider a double cover $`\pi :SP^2`$ of the projective plane, ramified along a smooth sextic $`CP^2`$ ($`S`$ is then realized as a K3 surface). Since any line in $`P^2`$ will intersect generically the sextic $`C`$ in six distinct point, we have that the covering $`\pi :SP^2`$ determines a (flat) family of hyperelliptic curves over the dual projective plane $`f:𝒳P^2`$. Then the Altmann-Kleiman compactification of the relative Jacobian of the family turns out to be a simplectic projective irreducible 4-folds, fibered over $`P^2`$, and in fact all fibres are Lagrangian Abelian varieties. Finally, we believe that our characterization of special Lagrangian submanifolds of irreducible symplectic 4-folds can be extended also to higher dimensional irreducible symplectic manifolds: to this aim notice that the proof we have given becomes longer and longer, since one has to deal with new cases and subcases. It would be nice, instead, to find out a sort of inductive argument, which works for all dimensions.
warning/0001/astro-ph0001390.html
ar5iv
text
# Emission-Line Galaxy Surveys as Probes of the Spatial Distribution of Dwarf Galaxies. I. The University of Michigan Survey ## 1 Introduction Studies which improve our understanding of the dependence of galaxy clustering on global galaxian properties such as morphology, luminosity and star-formation activity, are essential for deducing the correct model of galaxy formation. There is some expectation that the underlying mass distribution is smoother than the distribution of light, since it is unlikely that formation processes have been efficient enough to produce a one-to-one correspondence between dark matter halos and visible galaxies. Currently, the extent to which the overall galaxy distribution is biased relative to the underlying mass distribution is highly debated. The physical mechanisms which determine the large-scale biasing relation are complex, and presumably involve an interplay of processes such as gas cooling, star formation, supernova feedback and merging (Dekel & Rees (1987)). These processes may have affected different subsets of galaxies to varying degrees, with the result that the subsets now exhibit differing levels of large-scale clustering. Models of galaxy formation may therefore be constrained by testing for galaxy biasing through the statistical analysis of variations in the relative spatial distribution of different types of objects. Early work investigated the relationship between galaxy morphology and density environments (Davis & Geller (1976); Dressler (1980)), and the observational fact that early-type galaxies are more clustered than late-type galaxies on small scales is now well-known. Since then, the wealth of data provided by major redshift surveys over the past decade (e.g., CfA2, Geller & Huchra (1989); SSRS2, da Costa et al. (1994), 1998; Stromlo-APM, Loveday et al. (1996); Las Campanas Redshift Survey, Shectman et al. (1996)), has allowed for the investigation of the dependence of galaxy clustering on other properties such as color (Tucker et. al. (1995)), surface brightness (e.g., Thuan et al. (1987), 1991; Bothun et al. (1986), 1993), luminosity (e.g., Eder et al. (1989); Salzer et al. (1990); Loveday et al. (1995); Willmer et al. (1998)) and star-formation activity (e.g., Salzer (1989); Rosenberg et al. (1994); Telles & Maddox (1999)). In this work, we examine the relative spatial distributions of sub-samples of the overall galaxy population defined by two of these parameters, star-formation activity and luminosity. Although it has been generally well-established that star-forming galaxies are preferentially found in lower density environments than quiescent ‘normal’ galaxies (Iovino et al. (1988); Salzer (1989); Rosenberg et al. (1994); Pustil’nik et al. (1994); Telles & Terevich (1995); Loveday, et al. (1999)), it is not yet clear whether a relationship between clustering and luminosity exists. While some analyses suggest that clustering and luminosity are uncorrelated (e.g., Eder et al. (1989); Binggeli et al. (1990); Thuan et al. (1991); Weinberg et al. (1991)), there is also evidence that dwarf galaxies are more weakly clustered than the population of brighter giants (Salzer et al. (1990); Santiago & da Costa (1990); Park et al. (1994); Loveday et al. (1995); Willmer et al. (1998)). Resolving this discrepancy is of particular importance in constraining galaxy formation scenarios that involve biasing (“biased galaxy formation;” see for example Dekel & Rees (1987); White et al. 1987a ; Kauffmann et al. (1997)), since the majority of these models predict a rise in clustering with increasing galaxian luminosity. Nearly all of the above studies which investigate the dependence of clustering on luminosity (except for Eder et al. (1989) and Loveday et al. (1995)) draw their samples of dwarfs from surveys which are limited in magnitude at $``$15.5. Consequently, the intrinsically faint galaxies used in these analyses are confined to a small, local volume – few dwarfs beyond the Local Supercluster are included. This makes it difficult to draw definitive conclusions about the relative spatial distribution of dwarf and giant galaxies since the samples which are compared probe vastly different depths. In this study, we take a different approach to examining the clustering properties of high and low-luminosity galaxies. We augment the depth of the typical dwarf sample by combining dwarf emission-line galaxies from the University of Michigan (UM) objective-prism survey (MacAlpine et al. (1977)-1981) with dwarfs from the magnitude-limited Updated Zwicky Catalogue (UZC) (Falco et al. (1999)). The spatial distribution of this population of objects is then analyzed relative to samples of giant galaxies from the UZC. It has been shown that objective-prism surveys which select objects on the basis of line-emission are extremely effective at detecting intrinsically faint galaxies and constitute some of the deepest available samples of dwarfs (Salzer (1989)). This ability of emission-line surveys to sample further down the luminosity function partially helps to overcome the constraints imposed by magnitude-limited galaxy samples, in which the relatively rare luminous galaxies are over-represented while the more common dwarf galaxies are grossly under-represented. We refer to this selection effect throughout the paper as the Malmquist effect <sup>1</sup><sup>1</sup>1Not to be confused with the term Malmquist bias, which has through common usage come to have a different meaning in extragalactic astronomy. (a.k.a. the Scott (1956) effect). We stress that the emission-line selected dwarf samples used in this study act as surrogates for samples of the overall dwarf galaxy population. This is out of necessity, since existing magnitude-limited redshift surveys do not detect dwarfs at distances far beyond the Local Supercluster. For example, a dwarf galaxy with M<sub>B</sub> = $``$17.0 will only be included in a m<sub>B</sub> = 15.5 magnitude-limited survey if its velocity is under 2400 km s<sup>-1</sup>. For such nearby objects redshift is not an accurate measure of distance due to local velocity perturbations. In order to study the dwarf galaxy population beyond the Local Supercluster, selection techniques that are sensitive to dwarf galaxies at greater distances must be used. As mentioned above, surveys for emission-line objects provide samples of dwarfs which extend well beyond the Local Supercluster. Of course, there is no guarantee that the spatial distribution of line-selected galaxies will faithfully reflect that of the overall dwarf population. This is an important caveat which must be kept in mind as one interprets this work. While it would be preferable to carry out this type of study using a more representative sample of dwarf galaxies (including, for example, low-surface-brightness systems), the line-selected UM catalog is one of the few available with a large sample of dwarf galaxies beyond the Local Supercluster for which redshifts exist. Nevertheless, we do have reason to believe that our results will not be completely dominated by selection effects. The primary concern in using ELG dwarfs as probes of the overall dwarf population is that their activity may be a consequence of their clustering environment. If the star-formation episodes which gives rise to strong emission lines are triggered by galaxy-galaxy interactions, then line-selected dwarfs will be found, on average, in higher-density environments than quiescent dwarfs. Thus, clustering statistics calculated for the overall dwarf population using ELG dwarf samples will be overestimated. Since previous analyses have indicated that these strong-lined dwarfs are in fact less clustered, this effect does not seem to be present in our sample. This is not proof that emission-line dwarfs truly reflect the spatial distribution of the overall field galaxy dwarf population, but it at least gives us some confidence that the results we obtain will not be entirely biased. One should also keep in mind that even if these starbursting dwarfs are proper tracers of the dwarf galaxy large-scale distribution, ELG dwarfs still only represent a small fraction of the overall dwarf population. That is, only those dwarf galaxies currently undergoing a significant star-formation episode will be found in the UM survey. Once this episode fades in these particular galaxies, they too will become undetectable using current redshift-survey techniques. The fraction of dwarfs currently bursting (which is related to the so called “duty cycle”) is highly uncertain, and most likely varies in a complicated way that depends on galaxian properties such as the gas mass fraction and the shape of the underlying mass distribution. We make no attempt to account for the large additional population of dwarfs that are at the distances probed by the UM dwarfs, but are missing from the analysis because they are currently in a quiescent phase. In addition to investigating the dependence of clustering on luminosity, we also re-examine the relative clustering properties of emission-line galaxies (ELGs) and normal (non-active) galaxies. By including ELGs in all five lists of the UM survey and using a substantially more complete comparison sample of normal galaxies, we improve upon the Salzer (1989) analysis of the UM galaxies, which only examined objects in the fourth and fifth lists. Further, we extend Salzer (1989) by computing correlation functions as well as nearest neighbor statistics. The remainder of the paper is organized as follows. In $`\mathrm{\S }`$II, we describe the surveys from which galaxy samples are drawn. Newly acquired spectroscopic data for ELGs in the Fall UM survey area (lists I-IV) are presented in $`\mathrm{\S }`$III. A qualitative analysis of the relative spatial distributions of the UM ELGs and normal UZC galaxies is given in $`\mathrm{\S }`$IV via the visual inspection of cone diagrams of each of the UM survey regions. The clustering properties of the UM ELGs and dwarfs are quantified in $`\mathrm{\S }`$V, which describes our nearest neighbor and correlation function analyses in detail, and presents the results of the calculations. $`\mathrm{\S }`$VI summarizes our results and interprets them in the context of triggering mechanisms for galaxy activity and biasing in galaxy formation models. Throughout this paper a value of 75 km s<sup>-1</sup> Mpc<sup>-1</sup> is adopted for H<sub>o</sub>. ## 2 Galaxy Samples Emission-line galaxy samples are obtained from the University of Michigan (UM) objective-prism survey (MacAlpine et al. (1977)-1981). The physical properties of the samples and selection biases of the survey are discussed in detail in Salzer et al. (1989a, b) and Salzer (1989), so only a short description will be given here. The UM survey was conducted with the 61-cm Curtis Schmidt telescope in combination with an objective prism at the Cerro Tololo Inter-American Observatory. Objects in this survey are imaged on Kodak IIIa-J emulsion photographic plates and are selected based on the presence of \[O III\]$`\lambda \lambda `$4959,5007 and/or \[O II\]$`\lambda \lambda `$3726,3729 emission. The two separate regions of sky probed by the UM survey define the volumes whose galaxian spatial distributions are analyzed in this study. The first region resides in the autumn sky and consists of survey lists I through IV of the UM catalogue. Collectively, the 411 deg.<sup>2</sup> covered by these lists form an irregular strip extending from 23<sup>h</sup>12<sup>m</sup> to 2<sup>h</sup>20<sup>m</sup> in right ascension and from -2$`.^{}`$5 to +7$`.^{}`$0 in declination. The second region is located in the spring sky and is covered by UM list V. This region encompasses a rectangular 225 deg.<sup>2</sup> area ranging from 11<sup>h</sup>15<sup>m</sup> to 14<sup>h</sup>15<sup>m</sup> in right ascension and -2$`.^{}`$5 to +2$`.^{}`$5 in declination (see Figure Emission-Line Galaxy Surveys as Probes of the Spatial Distribution of Dwarf Galaxies. I. The University of Michigan Survey). Redshift data for the UM ELGs are compiled from Salzer et al. (1989a) and Terlevich et al. (1991). Also, we include previously unpublished spectroscopic data for an additional 26 UM survey objects, which are presented in the next section. Overall, the UM spring list provides a sample of 89 ELGs and the UM fall lists provide a sample of 179 ELGs. The comparison samples of normal (non-active) galaxies are composites of two separate galaxy catalogues. The new Updated Zwicky Catalogue (UZC) (Falco et al. (1999)), which is based upon the well-known Catalogue of Galaxies and Clusters of Galaxies (CGCG) (Zwicky et al. 1961- (1968)), provides a sample of objects with redshifts that is 98$`\%`$ complete to $`m_{Zw}`$15.5 for the right ascension ranges 20$`{}_{}{}^{\mathrm{h}}\alpha `$ 4<sup>h</sup>, 8$`{}_{}{}^{\mathrm{h}}\alpha `$ 17<sup>h</sup> and declination range -2$`.^{}`$5 $`\delta `$ 50. An unpublished compilation of galaxies provided by Giovanelli and Haynes supplies objects south of the CGCG survey limit. The latter catalogue is a compilation of galaxies from several published lists (e.g., CGCG and Uppsala General Catalogue of Galaxies (UGC, Nilson (1973)) in the north and the Morphological Catalogue of Galaxies (MGC, Vorontsov-Velyaminov et al. 1962- (1968)) and the ESO/Uppsala Survey of the ESO(B) Atlas (Lauberts (1982)) in the south). Maps of each of the UM analysis regions are shown in Figure Emission-Line Galaxy Surveys as Probes of the Spatial Distribution of Dwarf Galaxies. I. The University of Michigan Survey, where the sky positions of the UM ELGs and normal galaxies are indicated and the UM survey boundaries are marked with dashed lines. Note that a few UM galaxies are located outside of the nominal survey boundaries. Presumably this is due to slight variations in telescope pointing during the course of the UM survey. The regions covered by the objects in the comparison sample are extended relative to the UM survey regions. The ranges of these samples are 23<sup>h</sup>0$`{}_{}{}^{\mathrm{m}}\alpha `$ 2<sup>h</sup>36<sup>m</sup> and -5$`.^{}`$5 $`\delta `$ 10$`.^{}`$0 for the fall UM region, and 11<sup>h</sup>0$`{}_{}{}^{\mathrm{m}}\alpha `$14<sup>h</sup>30<sup>m</sup> and -5$`.^{}`$5 $`\delta `$ +5$`.^{}`$5 for the spring UM region. The comparison samples are extended to minimize errors due to boundary effects in the calculation of the nearest neighbor and correlation function statistics. This point will be discussed further in $`\mathrm{\S }`$5.2. The apparent and absolute magnitude distributions of the ELGs and UZC comparison galaxies in both of the UM survey areas are shown in Figure Emission-Line Galaxy Surveys as Probes of the Spatial Distribution of Dwarf Galaxies. I. The University of Michigan Survey. Clearly, the UM survey samples further down the luminosity function than the UZC. In fact, the ELG distributions are more heavily weighted towards the regime in which the magnitude-limited UZC becomes ineffective at detecting objects. Whereas the median absolute magnitudes of the UZC samples are -19.52 and -19.36 for the spring and fall regions respectively, the corresponding values for the UM ELG samples are -17.74 and -17.77. This increased depth is due to the selection of objects by line-emission. Because the equivalent-widths of the targeted lines, \[OIII\]$`\lambda \lambda `$4959,5007 and/or \[OII\]$`\lambda \lambda `$3726,3729, generally become larger as the luminosity of the host galaxy decreases, this technique is extremely effective at detecting intrinsically faint objects and partially helps to overcome the preferential selection of luminous galaxies at large distances (the Malmquist effect) in surveys which are limited by magnitude. The UM survey is thus an excellent source from which to draw samples of low-luminosity galaxies. We exploit this feature of the UM sample to study the relative clustering properties of low- and high-luminosity galaxies. ## 3 Spectroscopic Data Two major previous studies of the UM ELGs have obtained follow-up spectra for most of the 348 galaxy candidates in the survey (Salzer, MacAlpine & Boroson 1989a, b; Terlevich et al. 1991). As of the summer of 1993, redshifts were available, either from these sources or the literature, for all but roughly 60 of the UM galaxies. In an effort to acquire redshifts for as many of the remaining ELGs as possible, we obtained new optical spectra for 26 UM objects in July and August, 1993. All observations were obtained with the KPNO 2.1-m telescope using the Goldcam spectrograph <sup>2</sup><sup>2</sup>2Observations obtained at Kitt Peak National Observatory. KPNO is operated by AURA, Inc. under contract to the National Science Foundation. . The data were obtained as part of an experimental queue observing program being carried out at that time by NOAO. The detector used was a Ford 3072 $`\times `$ 1024 CCD with 15 micron pixels. Due to problems with the spectrograph camera optics, only approximately 2000 pixels were in reasonable focus, limiting the spectral range to 3600 – 6800 Å(at 1.49 Å/pixel). The grating used had 500 grooves/mm and was blazed at 5500 Å. A slit width of 2 arcsec was used for all observations. Reductions followed standard procedures (e.g., Salzer et al. 1995). Of the 26 ELG candidates observed, two were found to be QSOs (UM 57 & 230), one was a galaxy displaying only absorption lines at z = 0.102 (UM 23), five were galactic stars (UM 39, 44, 76,174, & 292), and 18 were emission line galaxies. The observational results for the ELGs are summarized in Table 1. The magnitudes listed in column 4 of the table come directly from the UM survey lists (MacAlpine et al. 1977-1981), and should only be taken as approximate. Column 6 lists the equivalent width of \[OIII\] $`\lambda `$5007, while columns 7 – 9 give the logarithms of the ratio of the emission-line fluxes for \[OIII\]/H$`\beta `$, \[OII\]/H$`\beta `$, and, when available, \[NII\]/H$`\alpha `$. Column 10 lists the Balmer decrement reddening parameter c. Note that for several of the galaxies, the H$`\alpha `$ and/or \[NII\] lines have been redshifted out of the available spectral range. Since c was not uniformly available, the line ratios in the table have not been corrected for reddening. All 18 ELGs observed here are starburst galaxies; none are Seyferts. With the addition of these new spectral data, only 32 of the 348 UM ELG candidates currently lack follow-up spectra. Most of these are quite faint (m$`{}_{B}{}^{}>`$ 18.0). ## 4 Qualitative Features of the Spatial Distribution The sets of cone diagrams shown in Figures Emission-Line Galaxy Surveys as Probes of the Spatial Distribution of Dwarf Galaxies. I. The University of Michigan Survey and Emission-Line Galaxy Surveys as Probes of the Spatial Distribution of Dwarf Galaxies. I. The University of Michigan Survey schematically present the spatial distributions of the galaxies in the fall and spring survey areas, respectively. Each set consists of two diagrams which plot (a) galaxies listed in the UZC alone and (b) galaxies in both the UM and UZC catalogues for recessional velocities between 0 and 10,000 km s<sup>-1</sup>. All radial velocities are corrected for the Galaxy’s motion with respect to the velocity centroid of the Local Group. In addition, galaxies which have been identified as residing in clusters or groups have had their radial velocities corrected for the velocity dispersion of the cluster. Table 2 lists all clusters/groups for which velocity dispersion corrections have been applied, along with the central coordinates, radii, cluster mean velocity, and velocity dispersion. These were identified by visual inspection of the cone diagrams, and the values of the parameters listed have been determined from the data in our galaxy catalog. Any group showing obvious velocity dispersion in the radial direction (“fingers of God”) is treated as a cluster. Comparison of the groups listed in Table 2 with previous studies of group/cluster identifications (e.g., Geller & Huchra 1983) shows good correspondence for the nearer groups, but our list includes a number of more distant clusters that were not recognized in the older redshift survey data. Cluster membership is determined on a galaxy-by-galaxy basis and depends on both spatial proximity to the group center (i.e., the galaxy lies within the cluster radius), as well as close velocity correspondence between each galaxy and the mean group velocity (observed velocity within $`\pm `$3$`\sigma `$ of the cluster mean). Once a galaxy is identified as a cluster member, its radial position in the cone diagram is modified in the following manner. If the observed radial velocity of the galaxy is $`\pm n\sigma `$ from the cluster mean, where $`\sigma `$ is the observed cluster velocity dispersion and n varies between 0 and 3, then the corrected velocity is $`\pm nrH_o/3`$ from the cluster mean velocity, where $`r`$ is the observed radius of the cluster (in Mpc) and $`H_o`$ is Hubble’s constant. This effectively collapses the observed velocity ellipsoid of a cluster down to a sphere. The radius $`r`$ is determined from the data and its value is chosen to include all galaxies that fall within the $`\pm `$3$`\sigma `$ velocity ellipsoid. In the cone diagram of UZC galaxies in the UM fall area (Figure Emission-Line Galaxy Surveys as Probes of the Spatial Distribution of Dwarf Galaxies. I. The University of Michigan Surveya), the most noticeable feature is the large band of galaxies that stretches laterally across the field at $``$5500 km s<sup>-1</sup>. The band begins as a narrow strip at the western edge of the plot and widens as it extends toward the east. This structure is part of the Pisces-Perseus Supercluster (Haynes & Giovanelli (1986)), the large sheet-like ensemble of galaxies which extends across nearly half of the South Galactic Cap. Five clusters can be found within the structure of the band, the richest of which is Abell 194 at $``$1$`\stackrel{\mathrm{h}}{\mathrm{.}}`$2. Toward the eastern edge of the plot at $``$2<sup>h</sup>15<sup>m</sup>, the broad end of the band forms the boundary of a void. In cone diagrams which cover a more extensive volume (Falco et al. (1999)), this void (diameter $``$1200 km s<sup>-1</sup>) can be seen to persist northward to a declination of $``$ 24 degrees. Two other voids of comparable extent inhabit the foreground. One underpopulated region, roughly centered with respect to the right ascension of the plot, is evident from 2500 km s<sup>-1</sup> to 4000 km s<sup>-1</sup> and extends to $`\delta `$ $``$48 degrees. One UZC galaxy, 00422+0453 (M<sub>B</sub> = -17.8) appears in the center of this void. Immediately behind this region in the western portion of the diagram lies the second void, which begins at 4000 km s<sup>-1</sup> and extends to the anterior edge of the large band. This void extends northward to a declination of at least 24 degrees. Behind the band, a larger underpopulated volume is evident from 6000 km s<sup>-1</sup> to 8000 km s<sup>-1</sup> and from 0<sup>h</sup> to 2<sup>h</sup> in right ascension. Five UZC galaxies, three of which are also detected by the UM survey, appear in this region. These galaxies are 00203+0633, 00279+0536, I0073 (UM 84), 01090+0104 (UM 307), and 01189+0322 (UM 96) (M<sub>B</sub> = -19.7, -19.7, -19.2, -20.4, -19.3 respectively). The cone diagram which plots the UM ELGs and UZC galaxies together (Figure Emission-Line Galaxy Surveys as Probes of the Spatial Distribution of Dwarf Galaxies. I. The University of Michigan Surveyc) reveals that there are an additional 7 ELGs (UM 47, 65, 98, 99, 119, 215, 92; M<sub>B</sub> = -18.7, -18.1, -18.4, -17.9, -17.1, -17.3, -16.3) which reside in this low-density area. These galaxies appear to be creating substructure, carving the low-density volume into smaller voids. A comparison of Figures Emission-Line Galaxy Surveys as Probes of the Spatial Distribution of Dwarf Galaxies. I. The University of Michigan Surveya and Emission-Line Galaxy Surveys as Probes of the Spatial Distribution of Dwarf Galaxies. I. The University of Michigan Surveyb show that the UM ELGs are concentrated in the areas where the UZC galaxies have formed structures - that is, the ELGs generally describe the same large-scale features as galaxies in the UZC. Clearly however, the ELGs are considerably more dispersed and avoid regions of high galaxian density. No UM ELGs appear in the clusters defined by the UZC galaxies. This latter observation is consistent with the findings of Dressler et al. (1985), Teague (1988) and Salzer (1989), and is likely due to gas-stripping mechanisms in high density clusters which inhibit activity. Turning now to the cone diagram of the UZC galaxies in the UM spring region, (Figure Emission-Line Galaxy Surveys as Probes of the Spatial Distribution of Dwarf Galaxies. I. The University of Michigan Surveya), the Local Supercluster (LS) is seen to dominate the region from 0 - 2000 km s<sup>-1</sup>. A filamentary structure extends outward from the LS to $``$3000 km s<sup>-1</sup> to form the boundary between two voids, one region in the east stretching from 1000 - 3000 km s<sup>-1</sup> which persists northward to $``$36 degrees, and another in the west from 2500 to 3500 km s<sup>-1</sup>. Three UZC galaxies, 11429+017, 11494-0222 and 12356+0014 (M<sub>B</sub> = -18.2, -18.1, -18.5 respectively), reside in this latter void. Figure Emission-Line Galaxy Surveys as Probes of the Spatial Distribution of Dwarf Galaxies. I. The University of Michigan Surveyb, where the UM ELGs and UZC galaxies are plotted together, reveals that there are an additional three ELGs, UM 454, UM 455 and UM 513 (M<sub>B</sub> = -16.9, -16.3, -16.2 respectively) in this void. A honeycomb-like structure appears in the region from 5000 - 7500 km s<sup>-1</sup>. The cluster MKW4 (Morgan, Kaiser & White (1975)) at $``$6000 km s<sup>-1</sup> extrudes from the eastern edge of the central ring of the honeycomb. Filamentary structures extend outward in all directions from this ring, connect with other filaments and define the neighboring voids. One UM ELG, UM 507 (M<sub>B</sub> = -16.2) appears within the void defined by the central ring of the honeycomb. Again, a comparison of Figures Emission-Line Galaxy Surveys as Probes of the Spatial Distribution of Dwarf Galaxies. I. The University of Michigan Surveya and Emission-Line Galaxy Surveys as Probes of the Spatial Distribution of Dwarf Galaxies. I. The University of Michigan Surveyb show that ELGs are more dispersed and avoid the major cluster (MKW 4) in this field. Other clusters identified in this field are listed in Table 2. There are a number of ELGs that reside in the neighborhood of the Local Supercluster. Nevertheless, these ELGs do not penetrate the densest regions of the LS – they only populate the periphery. The reader is referred to Salzer (1989) for an in-depth discussion of the ELG spatial distribution in this field. In summary, the qualitative analysis performed by visual inspection of these cone diagrams suggests that although the ELGs and normal galaxies outline the same large-scale features, the ELGs are less tightly confined to the features and tend to prefer regions of lower density. That is, compared to the population of normal galaxies, the ELGs appear to be less clustered. This suggests that a sub-sample of low-luminosity galaxies drawn from the UM survey would also be less clustered when compared to higher luminosity galaxies in the UZC. The statistical analyses conducted in the next section will quantify these results. In Figure Emission-Line Galaxy Surveys as Probes of the Spatial Distribution of Dwarf Galaxies. I. The University of Michigan Survey both of the UM fields have been plotted in a recessional velocity range extended to 20,000 km s<sup>-1</sup>. Examination of these diagrams shows a large difference in the depths of the samples. Figure Emission-Line Galaxy Surveys as Probes of the Spatial Distribution of Dwarf Galaxies. I. The University of Michigan Surveya shows that the normal galaxies in the UM fall region are well-represented out to a velocity of $``$12,000 km s<sup>-1</sup>. At this velocity, structures are still relatively well-defined, and a ridge of galaxies can be seen to cross the field, enclosing two voids in the foreground from 9000 to 12000 km s<sup>-1</sup>. Beyond 12,000 km s<sup>-1</sup> the normal galaxies become poorly sampled. Inspection of the normal galaxies in the UM spring region (Figure Emission-Line Galaxy Surveys as Probes of the Spatial Distribution of Dwarf Galaxies. I. The University of Michigan Surveyb), shows that this decline occurs at $``$10,000 km s<sup>-1</sup>. Beyond the outer boundary of the well-delineated honeycomb structure centered at $``$8000 km s<sup>-1</sup>, the normal galaxies become poorly sampled. The ELGs also suffer a decline in density with increasing distance. The rate of decline, however, is not as severe. A simple number count of galaxies in the fall region within the velocity range from 16,000 to 20,000 km s<sup>-1</sup> clearly illustrates this point – whereas, 10.2$`\%`$ (18/176) of the ELGs reside in this region, a mere 1.4$`\%`$ (7/495) of the normal galaxy sample can be found at these distances. In the spring UM region, the same effect is seen although there are fewer ELGs in this volume: 5.6$`\%`$ (5/89) of the ELG sample exists beyond 16,000 km s<sup>-1</sup> compared to only 0.3$`\%`$ (1/375) of the normal galaxy sample. These statistics attest to the power of using ELGs to chart large-scale structure in redshift regimes where magnitude-limited surveys are ineffective at detecting objects. Moreover, it is clear that it is not necessary to map every galaxy to gauge the general profile of the large-scale distribution. This combination of depth and efficiency indicates that there is great potential in using ELG samples and similar “sparse sampling” techniques to trace galaxian structure at ever larger distances. ## 5 Statistics of the Galaxy Distribution ### 5.1 Sample Construction The manner in which galaxy samples are defined is of particular importance to this analysis. Calculations are done on four groups of samples to deconvolve effects caused by the selection of ELGs as probes of the galaxian spatial distribution. This section highlights the important qualities of the various samples and attempts to provide a framework which the reader may find useful to refer to in later discussions of the statistical analyses (see in particular, Table 4). For each of the two regions considered in this study, three samples are constructed. The ELGs taken from the UM survey comprise one sample. UZC galaxies which reside within the areas covered by the UM survey form a sample of “normal” galaxies. Finally, UZC galaxies and galaxies from the unpublished catalogue provided by Haynes and Giovanelli in the extended regions (defined in $`\mathrm{\S }`$2) are taken as a computational comparison sample. The need for this latter sample will be made clear in $`\mathrm{\S }`$5.2. UM ELGs which appear in the normal galaxy and computational comparison samples are deleted from them and appear in the ELG sample only. Further, we exclude from the ELG sample a handful of objects in which the emission identified in the UM survey is due to normal disk HII regions (e.g., UM 505 and 506 from list V). Galaxies outside the velocity range from 2000 km s<sup>-1</sup> to 8000 km s<sup>-1</sup> are removed from the ELG and normal galaxy samples for the statistical analyses carried out below. The lower bound is necessary because radial velocity is not a reliable indicator of distance for near-by objects since peculiar velocities can be comparable to recessional velocities in the local neighborhood. Also, boundary effects on the computation of clustering statistics can be severe at the apex of the cone diagram. The upper bound is required to restrict the volume to well-sampled regions. Its determination is based on the discussion of sample depth in the previous section. As noted there, a rapid decline in the number of UZC galaxies occurs beyond a velocity of $``$10,000 km s<sup>-1</sup> in the fall region, and beyond $``$8000 km s<sup>-1</sup> in the spring region, while the UM ELGs are well-sampled beyond these velocities. Therefore, a conservative upper velocity limit of 8000 km s<sup>-1</sup> for the UM and UZC samples is imposed to avoid the “sampling mismatch” that can potentially produce false results in the following statistical analyses. At this velocity, an apparent magnitude limit of m<sub>Zw</sub>= 15.5 corresponds to a complete sampling of normal galaxies brighter than -19.6 (using H<sub>o</sub> = 75 km s<sup>-1</sup> Mpc<sup>-1</sup>). This upper bound of 8000 km s<sup>-1</sup> thus represents a compromise between the need to retain the lower luminosity galaxies and the desire to sample the largest volume possible. Since the computational comparison sample is meant to extend the limits of the normal galaxy sample to avoid boundary effects, no velocity restrictions have been imposed on its members. These three samples, the ELG and normal galaxies from 2000 km s<sup>-1</sup> to 8000 km s<sup>-1</sup> and the computational comparison sample in its entirety, represent one group of data with which the nearest neighbor and correlation function statistics are calculated. It is also necessary to perform these analyses on samples which consist of only field galaxies - galaxies which do not reside in gravitationally-bound clusters. This is to ensure that any statistical differences found to exist between the ELG and normal galaxy spatial distributions are not due to the paucity of ELGs in dense, clustered regions (as discussed in $`\mathrm{\S }`$4). Removing galaxies which reside in clusters from consideration also improves the consistency of the morphological makeup between the samples and decreases the likelihood that differences in clustering strengths are due to the morphology-density relation (Davis & Geller (1976); Dressler (1980)). This relationship suggests that elliptical galaxies preferentially reside in regions of high local galaxy density while spiral galaxies favor regions of low galaxy density. In particular, ellipticals appear to be largely confined to the virialized cores of clusters. Since the ELG samples are dominated by later-type galaxies (spirals and irregulars), removing cluster members from both the ELG and normal galaxy samples helps ensure that all samples have similar morphological compositions. For these reasons, a second group of data, in which galaxies residing in clusters (specified in Table 2) are eliminated from each of the samples, is constructed and analyzed. To investigate the spatial distribution of low-mass galaxies, and test the biased galaxy formation model prediction that intrinsically faint, dwarf galaxies are less clustered than the larger more luminous galaxies, a third group of data is constructed in which a sub-sample of low-luminosity galaxies is drawn from the dwarf-rich UM ELG sample. Objects from both the ELG and normal galaxy samples which are fainter than M<sub>B</sub> = -18.0 are taken as a sample of dwarfs, while UZC galaxies that are more luminous than -18.0 are taken as a sample of giants. The division between low- and high-luminosity galaxies is made by considering the desirability of including only the most extreme dwarfs in the analysis and the competing need to include a large enough number of objects in this sample from which reliable statistics may be calculated. As before, these samples are restricted to a velocity range of 2000 km s<sup>-1</sup> to 8000 km s<sup>-1</sup>, and a computational comparison sample without velocity limits is constructed from giant normal galaxies in the extended region. Since it is necessary to deconvolve morphology-density effects in this analysis as well, a fourth group of data from which cluster galaxies are removed from the dwarf, giant and computational comparison samples, is constructed and examined. It should be noted that in the current analysis luminosity is used as a surrogate for mass. Although it would be preferable to separate our dwarf and giant samples by mass, this is impractical given the available data. While we do not believe this choice compromises our conclusions, the existence of low-surface brightness galaxies like Malin 1 (Bothun et al. 1987) with large masses and relatively low luminosities suggests that some caution should be exercised when considering these results. We point out, however, that since our dwarf sample consists primarily of low-luminosity bursting ELGs ($``$86% of the fall and $``$76% of the spring dwarf samples are composed of ELGs), the dwarfs which are considered in this analysis have inflated luminosities and reduced mass-to-light ratios. Thus, selecting by luminosity has the effect of enhancing the mass differences between dwarfs and giants, rather than blurring them. In the following analyses, statistical quantities are calculated for a “test” population relative to a “control” population and for a control population relative to itself. Table 4 provides a summary of the contents in the four groups of data created for each region and indicates the samples which are considered test populations and those which are considered control populations. ### 5.2 Nearest Neighbor Distributions Given a sample of objects, the distribution of nearest neighbor separations may be directly calculated by measuring the distances between a target object and all other objects in the sample, retaining the minimum value, and repeating this procedure for every object in the sample. One can imagine that objects near the sample boundaries may have nearest neighbors which reside outside of the sample itself, so that distributions which are entirely constructed with separations found from within the sample may be skewed to larger separations. This is the reason that the extended computational comparison sample - which is simply a sample of the control population over a larger area - is introduced. In all of the cases below, the nearest neighbors of galaxies in the sample being considered are found in the extended computational comparison sample. This allows the periphery of the sample volume to be searched for nearest neighbors and prevents sample boundary effects from introducing errors into the analysis. For each group of data specified in Table 4, three different distributions of nearest neighbor separations are determined. The first distribution compiles “test object-control object” nearest neighbor distances and is the analog of the cross-correlation function computed in the next section. The second distribution compiles “control object-control object” nearest neighbor distances within the test volume and is the analog of the auto-correlation function, also computed in the next section. Comparison of these two distributions provides one measure of the difference in spatial distribution between the two samples. In particular, we apply Kolmogorov-Smirnov (KS) tests to the cumulative nearest neighbor distributions to determine the probability that the test galaxies and control galaxies are drawn from the same population of nearest neighbor separations. The third distribution provides a baseline for comparison and represents “random object-control object” nearest neighbor separations within the test volume. Calculations for 1000 simulated random samples of n galaxies, where n is equal to the number of objects in the test sample, are performed and averaged. The velocity distributions that are used to generate the random samples are determined by assuming a magnitude-limited (m$`{}_{Zw}{}^{}`$ 15.5) sample and a luminosity distribution consistent with the normal galaxies. For the latter, we utilize Schechter function parameters (Schechter 1976) derived from the normal galaxies in the computational comparison sample via the Lynden-Bell (1971) method. The calculated luminosity functions and Schechter function fits are plotted in Figures Emission-Line Galaxy Surveys as Probes of the Spatial Distribution of Dwarf Galaxies. I. The University of Michigan Surveya (fall region) and 6b (spring region). Values for M, $`\alpha `$ and $`\varphi ^{}`$ are indicated in the plots. Results from the nearest neighbor calculations are presented in Figures Emission-Line Galaxy Surveys as Probes of the Spatial Distribution of Dwarf Galaxies. I. The University of Michigan Survey (nearest neighbor separations for the ELG-normal galaxy data sets) and Emission-Line Galaxy Surveys as Probes of the Spatial Distribution of Dwarf Galaxies. I. The University of Michigan Survey (nearest neighbor separations for the dwarf-giant galaxy data sets). In all figures, histograms of the control and random samples have been scaled to the area under the test sample histogram. Plots for the samples in which all galaxies are included are shown in the left column of figures, while those for the samples from which cluster members are removed are shown on the right. The lower half of each figure shows the cumulative fraction distributions for the nearest neighbor distances plotted in the upper frame. Examination of Figure Emission-Line Galaxy Surveys as Probes of the Spatial Distribution of Dwarf Galaxies. I. The University of Michigan Surveya, which presents the results for the fall ELG/normal galaxy analysis where cluster members have been included, reveals that the ELGs are more dispersed than the normal galaxies. The normal galaxy histogram exhibits a large peak at small separations (between 0 and 1 Mpc) and a rapid decline in the number of objects beyond 2 Mpc. Very few objects are separated by larger distances. Only 9.1$`\%`$ (29/318) of the normal galaxies are seen to have nearest neighbor separations that are greater than 3 Mpc. In contrast, the ELG distribution has a much broader peak (between 1 Mpc and 3 Mpc), and 25.6$`\%`$ (23/90) of the ELGs have normal galaxy nearest neighbors which are greater than 3 Mpc away. The weaker clustering of the ELGs is perhaps most evident in the cumulative histogram diagram. The ELG curve rises more gradually than the normal galaxy curve, indicating that the ELGs are more dispersed. Nevertheless, it is also clear that the ELGs are more strongly clustered than a random distribution of galaxies. A KS test applied to the ELG and normal galaxy distribution shows that the hypothesis that the two samples have been drawn from the same population of nearest neighbor separations may be rejected at the 99.99$`\%`$ confidence level. Removing the cluster members from the samples has a clear impact on the nearest neighbor distribution of the normal galaxies, but minimally changes that of the ELGs. This reaffirms the observation that ELGs are not generally found in cluster environments. As would be expected, the large peak at small separations is eliminated in the normal galaxy histogram. At separations less than 3 Mpc the two distributions resemble each other more closely, but at larger separations they remain distinct. A KS test reveals that the differences in these distributions are significant at the 98$`\%`$ confidence level. The same trends may be discerned in the nearest distributions of the spring samples (Figure Emission-Line Galaxy Surveys as Probes of the Spatial Distribution of Dwarf Galaxies. I. The University of Michigan Surveyb). In the analysis which includes all galaxies, the ratio of ELGs to normal galaxies at large separations (greater than 3 Mpc) is approximately 2:1. This excess of large separations again decreases in the analysis from which cluster members have been removed, where the ratio is approximately 5:4 (see also Table 4). Elimination of the cluster galaxies does not have as great of an impact on the spring normal galaxy distribution because there are fewer clusters in the spring region. It has no effect on the ELG distribution. KS tests shows that the ELG and normal galaxy distributions are different at the 99.0$`\%`$ confidence level for the samples in which all galaxies have been included, and 91.3$`\%`$ for the samples from which cluster members are removed. The results of the nearest neighbor analyses for both fall and spring regions are consistent and show that the population of ELGs is less clustered than the population of normal galaxies. Turning now to the nearest neighbor distributions for the dwarf/giant galaxy samples (Figure Emission-Line Galaxy Surveys as Probes of the Spatial Distribution of Dwarf Galaxies. I. The University of Michigan Survey), it is important to bear in mind that the more pertinent analyses to examine are those in which cluster members have been excluded. This is because the samples of ELGs are now being used to trace the population of dwarf galaxies and are no longer being investigated in and of themselves. Thus, deconvolving selection effects caused by using the ELG population as the tracer is critical. Plots for the samples in which all galaxies are included are shown only for the purposes of comparison. In the fall analysis where cluster members have been removed (Figure Emission-Line Galaxy Surveys as Probes of the Spatial Distribution of Dwarf Galaxies. I. The University of Michigan Surveya), there is an excess of dwarf nearest neighbor separations beyond 3 Mpc. Whereas 26.6$`\%`$ (17/64) of the dwarfs have nearest neighbor separations greater than 3 Mpc, only 14.3$`\%`$ (33/231) of the normal galaxies are found in this regime. A KS test applied to the fall dwarf and giant distributions in which cluster galaxies have been removed shows that the two samples differ at the 99.1$`\%`$ confidence level. In the spring dwarf/giant samples in which cluster galaxies have been removed (Figure Emission-Line Galaxy Surveys as Probes of the Spatial Distribution of Dwarf Galaxies. I. The University of Michigan Surveyb), there is also an excess of dwarfs at large separations. 27.3$`\%`$ (9/33) of the dwarfs compared to 15.0$`\%`$ (29/193) of the giants are found at separations greater than 3 Mpc. There is also a substantial difference between the distributions at intermediate separations (2 Mpc to 6 Mpc). The dwarf nearest neighbor histogram appears to be shifted to larger separations relative to the giant galaxy nearest neighbor histogram. There is only one dwarf galaxy with a nearest neighbor separation less than 1 Mpc. A KS test applied to these distributions indicates that the two samples differ at the 99.9$`\%`$ confidence level. The results of the nearest neighbor analyses in the spring field are consistent with those of the fall field. Both analyses show that intrinsically fainter, less luminous galaxies are less clustered than the more luminous giant galaxies. ### 5.3 Correlation Functions To provide another statistical measure of clustering, correlation functions are computed for each group of data. First, the spatial auto-correlation function is calculated for the control sample. This statistic complements the nearest neighbor distributions in which control object-control object separations are measured. Second, the cross-correlation function between the test sample galaxies and control-sample galaxies is determined. This statistic complements the nearest neighbor distributions in which test object-control object separations are measured. The differences in clustering between the two samples are quantified by the differences in the two correlation functions. The method of computing correlation functions employed in this study is based on the discussion in Kirshner, Oemler, and Schechter (1979) and has been used by a number of authors, including Eder et al. (1989) and Salzer et al. (1990). Operationally, the auto-correlation function is defined by the relation $$1+\xi (r)=\underset{i=1}{\overset{n}{}}\frac{\rho _i(r)}{\overline{\rho }}$$ where $`\rho _i(r)`$ is the galaxian density at a distance r from the $`i^{th}`$ galaxy, $`\overline{\rho }`$ is the mean density of the entire sample volume, and n is the number of galaxies in the sample. Following this prescription, for a given separation length r, spherical shells with radius r and width $`\delta r`$ (chosen to be 0.125 Mpc) are centered on each of the objects in the sample, and the galaxian densities for the shells $`\rho _i(r)`$ are computed and summed. This sum is divided by $`\overline{\rho }`$ (determined from the luminosity function for the region) and yields $`\xi (r)`$ as given by the above expression. The technique for computing the cross-correlation function is the same, except that the shells are centered on galaxies in the test sample and the shell densities are determined from the comparison sample. In other words, $`\rho (r)`$ is the density of objects in the comparison sample at a distance r from an object in the test sample. As in the nearest neighbor computations, where the computational comparison sample is used to search for nearest neighbors, the computational comparison sample is used here to compute shell densities. All densities are corrected for decreased sampling with distance using the method described by Postman and Geller (1984). Correlation functions for the galaxy samples are presented in Figure Emission-Line Galaxy Surveys as Probes of the Spatial Distribution of Dwarf Galaxies. I. The University of Michigan Survey (ELG/normal galaxy sets) and Figure Emission-Line Galaxy Surveys as Probes of the Spatial Distribution of Dwarf Galaxies. I. The University of Michigan Survey (dwarf/giant galaxy sets) where log $`\xi (r)`$ has been plotted as a function of log r. The open circles represent the results from the auto-correlation analysis and the filled circles represent those from the cross-correlation analysis. Error bars are estimated from the Poisson noise ($`\sigma _p`$) in each bin. The solid lines are linear least-squares fits to the data and represent power-law functions of the form $`\xi (r)=(r_o/r)^\gamma `$ where the values of $`\gamma `$ and $`r_o`$ are indicated in each plot. The fits are weighted such that the weight of each point is equal to $`1/\sigma _p`$ and the errors given for $`r_o`$ and $`\gamma `$ are associated with the uncertainties in the fit. First, in examining the results of the correlation function analyses, it is reassuring to note that the auto-correlation function parameters (derived by fits to the open circles), calculated for the samples in which all galaxies are included, are generally consistent with the canonical values of 1.8 for $`\gamma `$ and 6.67 Mpc (5h<sup>-1</sup> Mpc with H=75 km s<sup>-1</sup> Mpc<sup>-1</sup>) for $`r_o`$. This is true for both the ELG/normal galaxy analysis and the dwarf/giant galaxy analysis. In the correlation function analyses for the ELG/normal galaxy samples, the correlation functions of the ELGs in all four plots tend to lie below that of the normal galaxies. In previous studies (Salzer et al 1990, Rosenberg and Salzer 1994), the direct comparison of the auto- and cross-correlation functions are facilitated by the fact that the powers ($`\gamma `$) of the two functions are identical within their errors. This allows for the calculation of a ratio of correlation amplitudes $`\xi _{normal}/\xi _{ELG}`$ that is independent of $`r`$. However in this study, the values of $`\gamma `$ for the auto- and cross-correlation functions differ considerably, which makes a direct comparison of the correlation functions difficult. Nevertheless, it is clear that the value of $`\xi _{ELG}`$ tends to be less than the value of $`\xi _{normal}`$, which implies that the ELGs are less clustered than the population of normal galaxies. This is consistent with the results of the nearest neighbor tests and confirms the conclusion that the population of ELGs are less clustered than the population of normal galaxies. To measure the difference in clustering, the ratio of the correlation function amplitudes, $`\xi _{normal}(r)`$ to $`\xi _{ELG}(r)`$, may be calculated at some fixed value of r. From the nearest neighbor analyses, a separation of 3 Mpc appears to be the characteristic value at which the ELG and normal galaxy distributions diverge, so the ratio is evaluated at r = 3 Mpc. The results are listed in Table 4. In the samples where all galaxies are included, the corrlation function amplitude of the normal galaxies is approximately twice that of the ELGs. The ratio drops to about 1.5 when cluster members are removed from the samples. The results of the dwarf/giant galaxy correlation function analysis also corroborate the results of the nearest neighbor tests of those samples. In Figure 10 it is clear that the correlation functions of the dwarf galaxies tend to lie below that of the giant galaxies. The disparity between the auto- and cross-correlation functions is particularly striking in the spring region. For the analysis where cluster members have been removed, the correlation length of the giant galaxies is about 3 times larger than that of the dwarf galaxies; that is, the dwarfs in the spring region have much less clustering power than their giant galaxy neighbors. This is most likely due to a combination of the relatively small number of galaxy clusters and prevalence of large voids in the spring region. The ratios of correlation function amplitudes, $`\xi _{giant}`$(3 Mpc) to $`\xi _{dwarf}`$(3 Mpc) are also shown in Table 4. The values of the ratios for the UM Fall dwarf samples are comparable to those of the UM Fall ELGs. In the UM spring region however, the ratios are significantly larger, again reflecting the characteristics of the galaxian distribution in the region. Of course, the errors associated with the ratios in the spring region are large, so care must be taken when interpreting the magnitude of this result. ## 6 Summary and Discussion This study uses emission-line galaxies (ELGs) drawn from the University of Michigan (UM) objective-prism survey to probe the spatial distribution of the dwarf galaxy population. The UM survey provides two samples of ELGs which cover widely separated areas on the sky. For each region a sub-sample of dwarfs (M$`{}_{B}{}^{}`$ -18.0) is analyzed relative to a sub-sample of higher luminosity galaxies (M$`{}_{B}{}^{}<`$ -18.0) from the Updated Zwicky Catalogue (UZC) which reside in the same volume of space. The relative clustering properties of UM ELGs and UZC normal galaxies, regardless of luminosity, are also studied. All quantitative analyses are carried out in the velocity range from 2000 km s<sup>-1</sup> to 8000 km s<sup>-1</sup>. ### 6.1 Galaxian Clustering and Activity In a previous paper which examined the relative spatial distribution of the UM ELGs, Salzer (1989) found that although ELGs follow the structures defined by normal galaxies, they tend to be less clustered and avoid regions of high galactic density. Salzer (1989) also noted that some ELGs are found in regions otherwise devoid of normal galaxies. With our new spectroscopic observations, inclusion of 18 additional ELGs from UM survey lists I-III, and a significantly more complete comparison sample of normal galaxies, our qualitative analysis of the cone diagrams of both fall and spring UM regions confirms and extends these conclusions. Results from nearest neighbor and correlation function analyses also show that star-forming galaxies are preferentially found in lower density environments than non-active normal galaxies. Nearest neighbor analyses reveal that the ELG samples have greater percentages of nearest neighbor separations at large values, while correlation function calculations indicate higher values of $`\xi `$ for the normal galaxy auto-correlation function as compared to the ELG-normal galaxy cross-correlation function. The conclusion that ELGs are less clustered than the overall galaxy population may not be too surprising since ELGs tend to be of late morphological type. However, even when the morphological make-up of the ELG and normal galaxy samples are homogenized by the removal of cluster galaxies, both tests still show a statistically significant difference between the clustering strength of the two populations. Table 4 summarizes the results from the nearest neighbor and correlation function analyses for all of the samples. These results are consistent with other studies of the relative spatial distribution of ELGs. Iovino et al. (1988), Rosenberg et al. (1994) and Loveday et al. (1999) have done correlation function analyses on ELGs and find a low clustering amplitude for this population. Pustil’nik et al. (1994) show that about 15% of their ELGs sample drawn from the Second Byurakan Survey and Case Low-Dispersion Northern Sky Survey are located within voids, indicating that active galaxies are more likely to inhabit regions of low galactic density than quiescent galaxies. A particularly puzzling issue which stems from these analyses concerns the cause of activity in ELGs. A popular hypothesis is that galaxy-galaxy interactions are the triggers of wide-spread star formation. While it is true that some ELGs are observed to be interacting with close companions (e.g., Salzer et al. (1995)), the result that ELGs tend to be more isolated than quiescent galaxies seems to suggest that interactions cannot be the primary cause of activity. Of course, the current study has only cross-correlated ELGs with objects from surveys which are magnitude-limited at $``$15.5, so our conclusion is limited to interactions with bright galaxies ($`M^{}\pm `$ 2). Searches for objects in the vicinity of ELGs which have found a deficit of $`L>L^{}`$ galaxies within 1 Mpc (Campos-Aguilar & Moles (1991); Campos-Aguilar et al. (1993); Vilchez (1995); Pustil’nik et al. (1995); Telles & Terevich (1995)) are consistent with our nearest neighbor analyses, which show few ELG-normal galaxy pairs at small separations, and confirms that ELGs do not typically have giant companions. It has been proposed that the observed activity in ELGs may be caused by interactions with fainter dwarfs or HI clouds (Brinks (1990)). However, recent work by Telles & Maddox (1999) which has cross-correlated ELGs in the Spectrophotometric Catalogue of HII Galaxies (Terlevich et al. (1991)) - a sample which has selected the majority of its objects from the Tololo (Smith et al. (1976)) and the UM surveys - with faint field galaxies in the Automatic Plate Measuring Machine catalogues has concluded that the active galaxies in their sample are not preferentially associated with faint, low mass companions. This leads to the general conclusion that interactions with galaxies, luminous or faint, are not the primary cause of the activity in ELGs. It is not clear what may be inducing the star formation in these galaxies. Although a large body of work is available that discusses triggering mechanisms involving tidal interactions (see for example Kennicutt et al. (1998)), very little has been done to investigate the causes of global starburst activity in isolated galaxies. Clearly, more effort must be invested in exploring alternate triggering mechanisms if further progress is to be made in understanding the formation and evolution of these galaxies. ### 6.2 Galaxian Clustering and Luminosity As discussed in both $`\mathrm{\S }`$1 and $`\mathrm{\S }`$2, samples of ELGs do not suffer from the Malmquist effect as severely as magnitude-limited samples and have luminosity distributions which are more heavily weighted towards the faint end. Moreover the redshifts of ELGs are easily determined, making this population not only an efficient probe of the dwarf galaxy population at distances well beyond the Local Supercluster, but also an effective tracer of the overall galaxian large-scale structure at distances where magnitude-limited samples become poorly sampled. In our analyses of the relative spatial distribution of low- and high-luminosity galaxies, we find that faint dwarfs are not as strongly clustered as their giant counterparts. There is an excess of dwarf-giant galaxy nearest neighbor separations of about 14% at distances larger than 3 Mpc, and the ratio of the giant galaxy auto-correlation function to the dwarf-giant cross-correlation function is approximately 2. Again, we have calculated these quantities for samples which we have attempted to homogenize with respect to morphology by removing all cluster galaxies. Our results are consistent with the scheme of “natural bias” in galaxy formation models which has been discussed by many authors (e.g., White et al. 1987a ; Cole & Kaiser (1989)). These models speculate that galaxies have preferentially formed in the peaks of the initial density field. For an initial field that is Gaussian, the larger but rarer 3$`\sigma `$-4$`\sigma `$ peaks would have been more clustered than the smaller, but more common 1$`\sigma `$-2$`\sigma `$ peaks. Presumably, massive dark matter halos, which have emerged from the more extreme density fluctuations, have formed luminous giant galaxies, whereas low-mass halos, which have collapsed from the smaller peaks, have formed fainter, lower mass galaxies. This leads to the prediction that high luminosity galaxies should be more clustered than the dwarf galaxies, which are less biased and better tracers of the overall mass distribution. In particular, it is interesting to note the agreement of our calculations with the models of White et al. (1987a & b) which predict differences between the correlation function amplitudes of luminous and dwarf galaxies of a factor of approximately 1.5 to 2. Of course, it cannot be entirely ruled out that what is being detected is evidence that ELGs are less clustered than the population of normal galaxies, rather than evidence of biasing, since our dwarf samples are primarily composed of ELGs. The excess number of nearest neighbor separations at large values and higher correlation function amplitudes are also found when statistics are calculated for the ELG/normal galaxy samples. The differences in clustering strength for the dwarfs relative to the giants, however, are greater than the differences in that of the ELGs relative to the normal galaxies. In the UM spring sample where cluster members have been eliminated, the ratio of dwarf nearest neighbor separations greater than 3 Mpc to that of the giant galaxies is larger when compared with the ratio of ELG nearest neighbor separations greater than 3 Mpc to that of the normal galaxies. In the UM fall samples these are approximately the same. In all samples the ratio of correlation function amplitudes is larger for the dwarf/giant samples than for the ELG/normal galaxy samples. Probably then, the differences in clustering strength between the emission-line selected dwarfs and the normal galaxy giants incorporate the effects of biasing and the phenomenon of diminished clustering with activity in galaxies. Until the ELG activation mechanism and its possible dependence on environment is properly understood, however, the relative contribution of these two effects to the reduced clustering amplitudes of our dwarf ELG samples will not be conclusively known. Further studies with larger and deeper samples of ELGs will allow us to more nearly deconvolve the dependence of clustering on activity and luminosity. With larger samples, we will be able to compare the clustering of bursting and non-bursting galaxies in several luminosity bins. In an upcoming paper (Lee, Salzer & Gronwall (2000); Paper II), we analyze the relative spatial distribution of ELGs detected in the new, ultra-deep KPNO International Spectroscopic Survey (KISS) (Salzer et al. (2000)). KISS provides a sample of ELGs that is considerably deeper - up to two magnitudes fainter - than other existing surveys for active galaxies. The area surveyed by KISS overlaps the Century Survey (Geller et al. (1997)) which covers the central 1 region of the first slice of the CFA Redshift Survey (de Lapparent, Geller & Huchra (1986)). In Paper II, we undertake analyses which closely follow those performed in this study, and compare our results with other observational studies of the luminosity-density relation. A discussion of the results of the current paper in the context of calculations by other groups will also be taken up there. We gratefully acknowledge financial support for this research through an NSF Presidential Faculty Award to JJS (NSF-AST-9553020). Early stages of this work were partially supported by a Cottrell College Science Award from the Research Corporation, as well as by research funds provided by Wesleyan University, to whom we are also grateful. Thanks are due to the following staff members of Kitt Peak National Observatory for carrying out the queue observations on our behalf: Todd Boroson, Diane Harmer and Jim DeVeny. We also thank Chris Impey and the anonymous referee for suggestions and comments that have improved the quality of this paper.
warning/0001/astro-ph0001053.html
ar5iv
text
# On the FIR emission from intracluster dust ## 1 Introduction Hot gas in clusters can potentially be primordial in origin, heated adiabatically during the infall into the potential well, or it could alternatively be hot gas that was ejected from galaxies into the intracluster (IC) medium. Studies of X-ray line emission have revealed surprisingly high metallicities (0.4-0.5), suggesting that non-negligible or even a substantial fraction of the intracluster medium originated in galaxies. It has also been suggested that the ejected material could contain dust and that this dust could produce detectable FIR emission (Dwek et al. 1990). However, in the hot X-ray emitting plasma from the centre regions of clusters of galaxies the dust is quite efficiently sputtered and destroyed, on time scales of typically a few times 10$`^8`$yr. To produce a detectable infrared emission this dust has to be injected at a high rate at the present epoch. Several indirect estimates indicating the presence of a substantial amount of grains in the IC medium, sufficient to detect in emission in the infrared, have been made. These have been derived from extinction measurements (Zwicky 1962; Karachentsev & Lipovetskii 1969; Bogart & Wagoner 1973; Boyle et al. 1988; Romani & Maoz 1992), from the soft X-ray absorption measurements (Voit & Donahue 1995, Arnaud & Mushotzky 1998) or from the observed amount of gas, by assuming that all the gas has been continuously injected with the Galactic gas-to-dust ratio, Z$`{}_{\mathrm{d}}{}^{}=0.0075`$ since formation of the cluster (Dwek et al. 1990). However, it is not clear whether the current populations of elliptical and spiral galaxies seen in clusters are capable of supporting the required injection rate to achieve detectability of IR emission from the IC medium, or whether other sources might be required. Observational evidence for FIR emission associated with intracluster dust has generally been inconclusive (Wise et al. 1993). Recently Stickel et al. (1998) detected a colour excess in the central 0.2 Mpc radius of Coma cluster, from the FIR emission measured by the ISOPHOT C200 camera aboard ISO. This colour excess was interpreted as thermal emission from intracluster dust with a temperature slightly higher than in the galactic cirrus and in cluster galaxies. However Quillen et al. (1999) have argued that the measurement by Stickel et al. comes from cluster galaxies. They estimated that the galaxies in the central region of the Coma cluster would produce a surface brightness of 0.06 MJy/sr, which they take as being in agreement with the Stickel detection. Nevertheless, the Stickel et al. measurement represents only a lower limit and it was obtained after subtracting the contribution of foreground galactic cirrus and cluster galaxies, on the assumption that the latter emission would arise from cooler dust. In summary there are major theoretical and observational uncertainties concerning both the amount and source(s) of IC dust. Here we make predictions for intracluster dust emission taking into account a variety of potential sources of dust. We illustrate the calculations using data for the Virgo Cluster, which is close enough for any intracluster IR emission to be spatially distinguished from emission from constituent galaxies and has detailed X-ray information, allowing realistic calculations of grain heating and sputtering time scales to be made. Furthermore, there is detailed information on all the Virgo galaxy members (Binggeli et al. 1993), and there has even been a recent detection of an intergalactic star population (Ferguson et al. 1998), which can be considered as a further potential source of intracluster grains. Finally, the Virgo cluster is the best studied example of a dynamically young cluster, into which spiral galaxies are falling in from the field (Tully $`\&`$ Shaya 1984). This allows a hitherto unconsidered, but potentially dominant, source of IC grains to be addressed, namely grains embedded in an external subvirial intergalactic (IG) medium accreting onto the X-ray emitting intracluster medium. The plan of this paper is as follows: In Sect. 2 we give details of the calculation of the infrared emission from stochastically heated grains in the Virgo IC medium, for grain populations arising from a balance between steady state injection and sputtering. In Sect. 3 we estimate the dust injection rate from potential sources (elliptical and spiral galaxies and IC stars) currently seen in the core of the Virgo cluster. We show that these sources are unlikely to provide enough dust to be detected in the IC medium by current infrared observatories. In Sect. 4 the feasibility of accretion of grains injected into an external intergalactic medium from distant infalling spiral galaxies is discussed. For plausible subvirial gas inflows and physical conditions in the intergalactic medium it is concluded that once grains have been injected into the intergalactic medium, they will descend largely unmodified into the cluster, approximately comoving with the infalling IG medium, and accumulate over the infall timescale. In Sect. 5 we make detailed estimates for the efficiency of the grain injection into the IG medium by calculating grain sputtering in outflow models for quiescent and star-burst galaxies. We find that dust grains can survive in winds from quiescent (non-starburst) galaxies and that their size distribution is not significantly changed by sputtering. In Sect. 6 we estimate the grain accretion rate from the B-band luminosity of the infalling population of spiral galaxies, finding this to be the main source of grains entering the IC medium. Once the grains reach the hot IC medium they radiate in the infrared and are sputtered on timescales smaller than the infall timescale. Calculations of the spectrum of the IR/submillimeter radiation of the accreted grains are given both for grains in the outer regions of the diffuse X-ray plateau, and for grains directly brought into contact with the dense core region. Some implications of the results are discussed in Sect. 7. A distance of 18.2 Mpc to the Virgo cluster is assumed throughout. ## 2 Calculation of the infrared emission In low-density hot astrophysical plasmas a dust particle is predominantly heated stochastically by the ambient gas and undergoes temperature fluctuations (Gail & Sedlmayr 1975, Draine & Anderson 1985, Dwek 1986, Dwek & Arendt 1992). To calculate the temperature distribution for different grain sizes in the cluster we adopted the parameters of the IC gas from Böhringer et al. (1994), Nulsen & Böhringer (1995) and Schindler et al. (1999) based on ROSAT observations. The X-ray morphology of the cluster was found to be very similar to the structure in the galaxy distribution, with a major component around M87 and a smaller component around M49. A faint diffuse component was also found to trace the cluster out to a distance of $`45^{}`$ from M87 (also seen by the Ginga satellite; Takano et al. 1989). The diffuse emission is rather asymmetric, falling off more steeply to the western side of the cluster. Because of the irregular structure, previous authors have divided the inner regions of the cluster into separate spherically symmetrical components centred on M87, M86 and M49, modelling the X-ray emission of each assuming hydrostatic equilibrium. Thus 71$`\%`$ of the total X-ray luminosity originates from the M87 halo (out to $`1^{}`$ from M87) and 15$`\%`$ comes from the diffuse X-ray component extending to $`45^{}`$ from M87. To calculate grain heating and sputtering rates within the cluster core we adopt the deprojected radial density and temperature profiles of Nulsen & Böhringer (1995), who divided the inner part of the cluster into 38 concentric spherical shells. Their results are reproduced in Table 1 (where we have binned together shells for which the temperature is constant). The abundances were fixed at 0.45 solar (Koyama et al. 1991; referred to Cosmic values in Allen 1973). Some recent results based on ASCA observations (Matsumoto et al. 1996) indicate a higher temperature for the hot plasma in the inner 10 of the cluster. However, as no deprojected model is yet available for the new ASCA data, we will use the ROSAT results throughout this paper. To calculate grain heating and sputtering rates appropriate for grains injected at the perimeter of the diffuse intracluster medium we estimated the plasma density at a distance of $`45^{}`$ from M87 using the results of Böhringer et al. (1994). These authors quote the total spatially integrated flux of the diffuse component as seen by ROSAT to be 15$`\%`$ of the total X-ray flux. Assuming spherical symmetry and the broad band emissivity being bremsstrahlung dominated, the corresponding average gas density is given by: $`\mathrm{n}_1^2={\displaystyle \frac{\mathrm{F}_1}{\mathrm{F}_2}}{\displaystyle \frac{\mathrm{n}_\mathrm{i}^2\mathrm{V}_\mathrm{i}\mathrm{T}_\mathrm{i}^{1/2}}{\mathrm{V}_1\mathrm{T}_1^{1/2}}}`$ (1) where n<sub>1</sub>, V<sub>1</sub>, F<sub>1</sub> and T<sub>1</sub> are the density, volume, flux and temperature of the diffuse X-ray component of the cluster and n<sub>i</sub>, V<sub>i</sub>, F<sub>i</sub> and T<sub>i</sub> are the same quantities, associated with the spherical shells within the halo of M87 (see Table 1). Since the temperature remains constant outside the halo of M87 with a temperature of $``$3 x 10$`^7`$K (Böhringer et al. 1994), we obtain an average gas density for the diffuse component of $`4\times 10^5\mathrm{cm}^3`$. This is comparable with the gas density at the edge of the diffuse emission region which can be derived from Schindler et al. (1999), for a model where the gas is in hydrostatic equilibrium. In the hot intracluster medium the grains are heated due to inelastic collisions with electrons and ions. In characterising the dust properties we considered spherical “astronomical silicate” grains and heat capacities from Guha-thakurta & Draine (1989), derived as a fit to experimental results for SiO<sub>2</sub> and obsidian at temperatures $`10<\mathrm{T}<300`$ K (Leger et al. 1985), with a simple extrapolation for $`\mathrm{T}>300`$ K. In Sect. 6 we also considered graphite grains, with heat capacities taken from Dwek (1986). The absorption efficiencies Q<sub>ν</sub>(a) were taken from Laor & Draine (1993) with grain sizes in the size interval \[$`\mathrm{a}_{\mathrm{min}},\mathrm{a}_{\mathrm{max}}`$\] from 10 $`\mathrm{\AA }`$ to 0.25 $`\mu `$m. For the heating of the grains in a hot plasma we adopted the method used by Dwek (1987). Here the grain is assumed to have an effective thickness $`\mathrm{R}_0=4\mathrm{a}/3`$ and let R(E) be the range at which gas particles with kinetic energy E would be stopped. If R(E) is shorter than the effective thickness, then all energy will go onto the grain. Otherwise the gas particles will have a rest energy E’, that is given by $`\mathrm{R}(\mathrm{E}^{})=\mathrm{R}(\mathrm{E})\mathrm{R}_0`$. In the case of electron heating we used an analytical expression of the electron stopping power R(E)$`\rho `$ for silicate and graphite from Dwek & Smith (1996). In the case of ion heating the energy fraction deposited by ions onto the grains was calculated with a formula similar to that used by Dwek & Werner (1981), which is based on an approximation from Draine & Salpeter (1979), for the range of H and He in solids up to 100 keV: $`\zeta (\mathrm{E})=\{\begin{array}{cc}1\hfill & ,\mathrm{E}<\mathrm{E}^{},\hfill \\ \mathrm{E}^{}/\mathrm{E}\hfill & ,\mathrm{otherwise},\hfill \end{array}`$ (4) where E is the minimum kinetic energy of a nucleus to penetrate the grain. It is given by $`\mathrm{E}^{}[\mathrm{keV}]={\displaystyle \frac{1}{3}}\mathrm{a}[\mu m]\rho [\mathrm{g}/\mathrm{cm}^3]\times `$ (5) $`\{\begin{array}{cc}133\hfill & \mathrm{for}\mathrm{H}\mathrm{atoms},\hfill \\ 222\hfill & \mathrm{for}\mathrm{He}\mathrm{atoms},\hfill \\ 665\hfill & \mathrm{for}\mathrm{C},\mathrm{N},\mathrm{O}\mathrm{atoms}.\hfill \end{array}`$ (9) This shows that in a hot gas the ion heating of very small grains is nearly discrete. In low density plasma, like that existing in the intracluster medium, the grains are mainly stochastically heated. The stochastic heating processes are calculated following the method of Guhathakurta & Draine (1989). This method derives the temperature distribution P(a,T<sub>d</sub>) of various grain radii a as a function of dust temperature T<sub>d</sub>. In Fig. 1 we give the temperature distribution P(a,T<sub>d</sub>) for various silicate grain sizes, for the inner 1$`{}_{}{}^{}.67`$ radius of the Virgo Cluster (n$`{}_{\mathrm{e}}{}^{}=0.0358`$ $`\mathrm{cm}^3`$, T$`{}_{\mathrm{e}}{}^{}=1.30\times 10^7`$). Due to stochastic heating, small grains undergo significant fluctuations from the equilibrium temperature, while larger grains have probability functions becoming narrower, eventually approaching delta functions. Many of the smaller grains exhibit a plateau in their probability distribution, separated by steps corresponding to the energy thresholds at which the various ions are stopped by the grains (see Eqs. (2) and (3)). In order to calculate the contribution of all grain sizes we have to derive the grain size distribution. We consider that the grains are continuously injected in the IC medium with a grain size distribution given by a power law, a<sup>-k</sup>. The mechanism through which they can be injected is discussed in the next sections; here we consider the general case. The grain size distribution can be determined approximately by balancing the local rate of grain injection into the ICM, S(a), and the rate of grain destruction by sputtering: $`{\displaystyle \frac{}{\mathrm{a}}}\left(\mathrm{N}(\mathrm{a}){\displaystyle \frac{\mathrm{da}}{\mathrm{dt}}}\right)=\mathrm{S}(\mathrm{a})`$ (10) where N(a) da is the total number of particles with sizes in the interval \[a,a+da\]. For gas temperatures $`10^6<\mathrm{T}<10^9`$ K the sputtering timescale for a silicate or graphite dust particle of radius $`a`$ is given by Draine & Salpeter (1979): $`\mathrm{t}_{\mathrm{sput}}[\mathrm{yr}]{\displaystyle \frac{10^6\mathrm{a}[\mu \mathrm{m}]}{\mathrm{n}_\mathrm{H}[\mathrm{cm}^3]}}`$ (11) where n<sub>H</sub> is the gas density. From Eqs. (4) and (5) results the steady state grain size distribution: $`\mathrm{N}(\mathrm{a})={\displaystyle \frac{\mathrm{t}_{\mathrm{sput}}}{\mathrm{a}}}{\displaystyle \frac{\mathrm{dM}_\mathrm{i}}{\mathrm{dt}}}{\displaystyle \frac{\mathrm{a}^{k+1}\mathrm{a}_{\mathrm{max}}^{k+1}}{\mathrm{k}1}}\times `$ (12) $`\times \left[{\displaystyle _{a_{\mathrm{min}}}^{a_{\mathrm{max}}}}\mathrm{a}^k{\displaystyle \frac{4\pi \mathrm{a}^3}{3}}\rho \mathrm{da}\right]^1`$ where $`\rho `$ is the density of the dust grains and $`\mathrm{dM}_\mathrm{i}/\mathrm{dt}`$ is the total dust injection rate: $`{\displaystyle \frac{\mathrm{dM}_\mathrm{i}}{\mathrm{dt}}}={\displaystyle _{a_{\mathrm{min}}}^{a_{\mathrm{max}}}}\mathrm{S}(\mathrm{a}){\displaystyle \frac{4\pi \mathrm{a}^3}{3}}\rho \mathrm{da}`$ (13) The total dust mass is then given by: $`\mathrm{M}={\displaystyle _{a_{\mathrm{min}}}^{a_{\mathrm{max}}}}\mathrm{N}(\mathrm{a}){\displaystyle \frac{4\pi \mathrm{a}^3}{3}}\rho \mathrm{da}`$ (14) and the total infrared emission is: $`\mathrm{F}_\nu `$ $`=`$ $`{\displaystyle \frac{1}{\mathrm{d}^2}}{\displaystyle _{a_{\mathrm{min}}}^{a_{\mathrm{max}}}}\mathrm{N}(\mathrm{a})\mathrm{da}\pi \mathrm{a}^2\mathrm{Q}_\nu (\mathrm{a})\times `$ (15) $`\times {\displaystyle _0^{\mathrm{}}}\mathrm{B}_\nu (\mathrm{T}_\mathrm{d})\mathrm{P}(\mathrm{a},\mathrm{T}_\mathrm{d})\mathrm{dT}_\mathrm{d}`$ where B<sub>ν</sub> is the Planck function. ## 3 Predicted FIR emission from intracluster dust produced by sources inside the core region of the Virgo cluster (within $`1^{}`$ of M87) ### 3.1 Dust originating in individual intergalactic stars In this section we estimate the amount of dust produced through stellar winds by individual intergalactic red-giant-branch stars, as detected recently in the Virgo cluster by Ferguson et al. (1998). The motivation to consider these stars as potential sources of dust was the suggestion that the intergalactic stars are likely to have originated primarily from the elliptical and S0 galaxies, and thus to contain a higher proportion of M supergiants to giants as compared to the Galactic disk, and therefore a higher integrated mass-loss and dust production. Furthermore, this dust would not suffer sputtering losses in the injection process, as is the case for injection from galaxies into the intracluster medium through galactic winds. According to Ferguson et al. the relatively smooth distribution of mass inferred from the X-ray observations suggests that most of the intergalactic material was stripped by tidal interactions with the cluster potential. Since early-type galaxies are more numerous in the cluster, have older stellar populations, and are likely to have inhabited the central megaparsec of the cluster for much longer than the spiral and irregular galaxies, we will consider here that the intergalactic stars were stripped from early type galaxies. Nevertheless, the possibility that some of the stars formed in situ, or that some were stripped off by impulsive interactions between galaxies (‘harassment’, Moore et al. 1996) cannot be ruled out (Ferguson et al. 1998). The observations made with the Hubble Space Telescope WFPC2 camera with the F814W (approximately I-band) filter on a blank field in the Virgo cluster (Ferguson et al. 1998) indicated a clear excess in source counts in the Virgo cluster image. The total flux from the excess sources was used by Ferguson et al. to calculate the total mass of stars below their detection limits. For this they considered a population with a metallicity, expressed as a decimal logarithm relative to the solar iron-to-hydrogen ratio, \[Fe/H\]=-0.7, an age of 13 Gyr, a Salpeter initial mass function, and a distance of 18.2 Mpc. They derived an underlying surface mass density of 0.14 M$`\mathrm{pc}^2`$ if the initial mass function continues to 0.1 $`\mathrm{M}_{}`$. To calculate the total mass loss of all the stars within the X-ray emitting core region of 60 arcmin radius from the position of M87, we consider that the stars follow the distribution of elliptical (E) and S0 galaxies. Their distribution is described by a spherical King model: $`\rho (\xi )=\rho _0(1+\xi ^2)^{3/2}`$ (16) where $`\rho _0`$ is the central mass density of stars, $`\xi =\mathrm{r}/\mathrm{r}_\mathrm{c}`$ is the radius in units of the core radius $`\mathrm{r}_\mathrm{c}`$. We take the core radius $`\mathrm{r}_\mathrm{c}=1.1^{}`$, as given for the E+S0 Virgo galaxies by Binggeli et al. (1987). The projected mass surface density of stars is then: $`\mathrm{\Sigma }(\lambda )={\displaystyle \frac{2\rho _0\mathrm{r}_\mathrm{c}}{(1+\lambda ^2)}}\left(1{\displaystyle \frac{1+\lambda ^2}{1+\xi _\mathrm{t}^2}}\right)^{1/2}`$ (17) where $`\lambda =\mathrm{R}/\mathrm{r}_\mathrm{c}`$ is the projected radius in units of the core radius and $`\xi _\mathrm{t}=\mathrm{r}_\mathrm{t}/\mathrm{r}_\mathrm{c}`$ is the radius of the outer boundary in units of core radius. If we consider a surface mass density of intracluster stars of 0.14$`\mathrm{M}_{}\mathrm{pc}^2`$, as derived by Ferguson et al. (1998) at a distance of 44.5 arcmin from M87, from Eq. (10) we obtain the central mass density of stars, $`\rho _0=0.363\times 10^6\mathrm{M}_{}\mathrm{pc}^3`$. The stellar mass within radius $`\xi `$ in units of the stellar core mass M<sub>c</sub> is: $`\mathrm{M}(\xi )=3[\mathrm{ln}(\xi +(1+\xi ^2)^{1/2})\xi (1+\xi ^2)^{1/2}]`$ (18) with the core mass $`\mathrm{M}_\mathrm{c}={\displaystyle \frac{4\pi \mathrm{r}_\mathrm{c}^3\rho _0}{3}}`$ (19) From Eqs. (10) and (11) we calculate the stellar mass in each of the spherical shells from Table 1 and the results are given in Table 2, Column 2. In order to obtain the number of stars of different spectral types and luminosity classes that correspond to the derived total mass we have to integrate over the mass function of such a population. We consider a model HRD population for an elliptical galaxy assuming the same parameters as used by Ferguson et al. (1998) to derive the total mass of the stars. For the mass-loss rates and gas to dust ratio in the outflow of different types of stars we took the values from Whittet (1992) and Gehrz (1989) for the Galactic stellar population. Whilst bearing in mind that the averaged mass-loss and dust to gas ratio are subject to considerable uncertainty, we may draw some general conclusions from these data. Summing individual injection rates for stardust we obtain a total injection rate of dust $`\dot{\mathrm{M}}_\mathrm{i}`$ for each spherical shell, and the corresponding values are given in Table 2, Column 3. The main contribution to the injection rate comes from M supergiants, which would imply that most of the grains should consist of silicates. The values of the dust injection rate are quite low, indicating that there are not enough stars to produce a detectable amount of dust in the inner hot region of the cluster, since the grains are sputtered very efficiently by the ambient hot gas. We assume that the grain size distribution injected into the IC medium by the stars is given by a power law with an exponent given by the MRN (Mathis et al. 1977) value of k=3.5. We also assume a silicate dust composition. From Eqs. (8) and (9) we calculated the infrared intensity at different projected distances from the cluster by integrating over the line of sight through the cluster. The corresponding infrared spectrum in the inner 1$`{}_{}{}^{}.67`$ region of the Virgo cluster is plotted with a solid line in Fig. 2, while in Fig. 3 we give the radial brightness profile at 100 $`\mu `$m. The infrared emission from dust injected by intergalactic stars inside the core region of the Virgo cluster is less than 0.01 MJy/st, about an order of magnitude below the detection limit of $``$0.1 MJy/st for currently available telescopes. ### 3.2 Dust originating in early type galaxies The elliptical and S0 galaxies are not only the best candidates for the sources of the intracluster stars (discussed in the previous section), but they are also thought to release gas into the ICM via supernovae-driven galactic winds. Because these galaxies have remained in the inner region of the cluster since their formation, their integrated mass-loss has been used (Okazaki et al. 1993) to calculate the galaxian contribution to the observed amount of gas in the centre of the cluster. Okazaki et al. selected all E, S0 and dwarf elliptical galaxies from the Virgo Cluster Catalogue of Binggeli et al. (1985), in the field within 3 from M87. They used three models for elliptical galaxy formation with galactic winds (Arimoto & Yoshii 1987, Matteucci & Tornambe 1987, David et al. 1990,1991a,b) and estimated the masses of gas that have been ejected from all these galaxies during the cluster lifetime. The result was that even the largest mass obtained by one of their model prediction is only 10$`\%`$ of the value derived from X-ray observations. They concluded that elliptical galaxies cannot produce the observed amount of gas and that 90$`\%`$ of the gas must be of primordial origin. In the case of the dust released by elliptical galaxies via supernovae-driven galactic winds, it is obvious that only the dust produced quite recently can give rise to an IR emission, since the dust is quickly sputtered away. But elliptical galaxies have released most of their gas in the earlier epochs of the cluster lifetime, so their current injection rate is low. We will show that even for the most optimistic assumption, of a steady state mass-loss over the cluster lifetime, only a negligible amount of dust is predicted, since the elliptical galaxies are well depleted of dust in comparison with the spiral galaxies (Tsai & Mathews 1996). Thus if we take the amount of gas predicted by Okazaki et al., an average dust-to-gas ratio a hundred times lower than the Galactic value (Tsai & Mathews 1996) and if we divide this value by the cluster lifetime we obtain an average dust injection rate which is an upper limit for the present dust injection rate. Tsai & Mathews (1995) have shown that in the elliptical galaxies the grain size distribution is given by a power law with index $`(\mathrm{k}1)`$. They have also shown that this solution is valid when the sputtering time is short compared to the radial flow time, but the same solution is still an excellent approximation even when the sputtering time is so long that the grains can move inwards during their lifetime across an appreciable part of the galaxy. Assuming again that the grains in the elliptical galaxies are injected from the evolving stars with a power law having the MRN index $`\mathrm{k}=3.5`$, the size distribution of the ellipticals will be a power law of index 2.5. We also assume that the grain size distribution is not modified inside the galactic wind and that the dust consists of silicate grains. With these assumptions we calculated the dust injection rate in the same spherical model for the cluster core as previously used (Table 1). The results are given in Table 3. The corresponding spectrum and brightness profile at 100 $`\mu `$m are given in Figs. 2 and 3 with dashed lines. The amount of infrared emission produced by dust ejected by all early type galaxies in the inner core of Virgo cluster is negligible even in the upper limit calculation, being only at best comparable to the emission from stellar dust. ### 3.3 Dust originating in late-type galaxies Late-type galaxies are known to contain more dust than the elliptical galaxies, and they can eject interstellar gas through interaction with the cluster environment. Haynes & Giovanelli (1986) have shown that Virgo spirals within 5 degrees of M87 are HI deficient compared with their field counterparts. The HI deficiency becomes more marked for the central regions, with three-quarters of Virgo spirals within 2.5 degrees of M87 deficient by more than a factor 3 ($`\mathrm{DEF}>+0.47`$). The summation of HI deficiency by gas mass for spirals seen in projection within 5 degrees of M87 (as tabulated by Giovanelli & Haynes 1983 and Haynes & Giovanelli 1986) is 5.7$`\times `$10$`^{10}`$M, with the bulk of the summed deficiency being due to the giant spirals. Taking the dust (and gas) replenishment timescale in galactic disks to be comparable to that estimated for the Milky Way of $``$3$`\times `$10<sup>9</sup>yr (e.g. Jones et al. 1997) one can estimate a current gas injection rate of $``$19 Myr<sup>-1</sup> into the Virgo ICM from spiral galaxies. The dust content of the diffuse HI prior to ejection is difficult to estimate. On the one hand gas is preferentially lost from the outer disk, where the metallicity and dust abundance is typically lower than that for the galaxy as a whole. On the other hand, it is known that spirals in the Virgo core have an enhanced metallicity by a factor of approximately 2 compared with field counterparts of similar lateness (Skillman et al. 1996). Assuming that the ejected gas originally contained dust with an abundance of 0.0075 by mass (i.e. that of the Milky Way), we obtain a crude upper limit for the dust injection rate into the Virgo ICM of 0.14 Myr<sup>-1</sup>. This upper limit corresponds to the case that grain sputtering in the injection process from the interstellar to the intracluster medium can be ignored. It is somewhat lower that the estimated injection rate from the intracluster star population. Whether the actual dust injection from spirals into the IC medium approaches this upper limit is likely to depend on the mechanism for ejection. The most favoured model for gas removal from the HI deficient galaxies, as originally proposed by Gunn & Gott (1972), is ram-pressure stripping, whereby gas can be removed if the ram pressure P$`{}_{\mathrm{ram}}{}^{}\rho _{\mathrm{ICM}}\mathrm{v}_{\mathrm{gal}}^2`$ exceeds the gravitational restoring force per unit area on the diffuse HI gas, P$`{}_{\mathrm{grav}}{}^{}=2\pi \mathrm{G}\sigma _{\mathrm{grav}}\sigma _{\mathrm{HI}}`$. Here v<sub>gal</sub> is the relative velocity of the galaxy through the IC medium of density $`\rho _{\mathrm{ICM}}`$, and $`\sigma _{\mathrm{grav}}`$, $`\sigma _{\mathrm{HI}}`$ are the gravitational and HI gas surface densities in the galactic disk, respectively. Typically, giant spirals have $`\sigma _{\mathrm{HI}}\mathrm{\hspace{0.17em}10}\mathrm{M}_{}\mathrm{pc}^2`$ averaged over the disk (e.g. Roberts & Haynes 1994; Young & Scoville 1991); the Milky Way has $`\sigma _{\mathrm{HI}}\mathrm{\hspace{0.17em}5}\mathrm{M}_{}\mathrm{pc}^2`$ at the solar circle (Dickey & Lockman 1990). Within 1 of M87, $`\rho _{\mathrm{ICM}}\mathrm{\hspace{0.17em}10}^{24}\mathrm{kg}\mathrm{m}^3`$, and typically v$`{}_{\mathrm{gal}}{}^{}\mathrm{\hspace{0.17em}1000}\mathrm{km}\mathrm{s}^1`$. This is sufficient to strip diffuse interstellar HI with $`\sigma _{\mathrm{HI}}\mathrm{\hspace{0.17em}10}\mathrm{M}_{}\mathrm{pc}^2`$ from within a giant spiral like the Milky Way, which has (Bahcall et al. 1992) $`\sigma _{\mathrm{grav}}\mathrm{\hspace{0.17em}70}\mathrm{M}_{}\mathrm{pc}^2`$ at the solar circle. To estimate whether grains would be sputtered in this process we note that a sudden interaction of a galaxy with the IC medium would generally drive a shock wave into a cold diffuse HI disk of density $`\rho _{\mathrm{HI}}`$ with speed v$`{}_{\mathrm{s}}{}^{}(\rho _{\mathrm{ICM}}/\rho _{\mathrm{HI}})^{0.5}\mathrm{v}_{\mathrm{gal}}`$. For mid-plane number densities of $`\mathrm{\hspace{0.17em}0.5}\mathrm{cm}^3`$, as in the diffuse HI in the plane of the Milky Way at the solar circle (Dickey & Lockman 1990), v<sub>s</sub> is 37$`\mathrm{km}\mathrm{s}^1`$, for v$`{}_{\mathrm{gal}}{}^{}\mathrm{\hspace{0.17em}1000}\mathrm{km}\mathrm{s}^1`$ and $`\rho _{\mathrm{ICM}}\mathrm{\hspace{0.17em}10}^{24}\mathrm{kg}\mathrm{m}^3`$ (i.e. 1 from M87). This shock speed is well below the minimum value of $`\mathrm{\hspace{0.17em}100}\mathrm{km}\mathrm{s}^1`$ needed for sputtering. Although sputtering should become effective in ram-pressure stripping from the outer disk, where HI number densities may be orders of magnitude lower, this simple consideration is consistent with the possibility that a substantial fraction of the interstellar grains may survive the ram-pressure stripping process. Indeed, some direct evidence for the presence of dust in interstellar material stripped from the Virgo elliptical galaxy M84 has been found by Rangarajan et al. (1995). High resolution hydrodynamical simulations of ram-pressure stripping of elliptical galaxies (Balsara et al. 1994) show that the gas removal involves a series of individual events separated by $`\tau _{\mathrm{strip}}`$ of a few times 10$`^7`$yr, leading to long tongues of stripped gas in the IC medium, a result which Balsara et al. consider would also apply to gas stripping from spirals. Once injected into the IC medium, the grains will be rapidly sputtered by the hot gas, so that the infrared emission should still be morphologically associated with the parent galaxy as a dust trail (Dwek et al. 1990), rather than being smoothly distributed in the IC medium. For the example of an ambient number density of 0.0007$`\mathrm{cm}^3`$ (corresponding to 1 from M87 - see Table 1), the sputtering time scale (Eq. (5)) is 1.4$`\times `$10$`^8`$yr for grains of size 0.1$`\mu m`$. The length of the infrared-emitting dust trail is then the distance traveled by the galaxy in this sputtering time, or 146 kpc for v$`{}_{\mathrm{gal}}{}^{}\mathrm{\hspace{0.17em}1000}`$km s<sup>-1</sup>. We have estimated the infrared emission from a giant spiral with a radius of 20 kpc and an initial HI mass of about $`6.6\times 10^9\mathrm{M}_{}`$. For a typical HI deficiency (Haynes & Giovanelli 1986) of DEF=+0.47, the total HI mass loss is $`4.4\times 10^9\mathrm{M}_{}`$. Assuming most of the loss is sudden, on entering the cluster core region, and taking the dust-to-gas ratio to be $`\mathrm{Z}_\mathrm{d}=0.0075`$ (the Galactic value), then the mass of the dust released in the ICM medium by such a galaxy is $`\mathrm{M}_d=3.3\times 10^7\mathrm{M}_{}`$. The infrared spectrum of such a source is given in Fig 4, for two cases; the first case (solid line) is when the stripping occurs at 10 from the cluster centre, ($`\mathrm{n}_\mathrm{e}=0.0039\mathrm{cm}^3`$, $`\mathrm{T}_\mathrm{e}=2.36\times 10^7`$ K, see Table 1) and the second case (dashed-line) is when stripping occurs while the galaxy enters the dense core region of the cluster, at $`1^{}`$ distance from the centre of the cluster ($`\mathrm{n}_\mathrm{e}=0.0007\mathrm{cm}^3`$, $`\mathrm{T}_\mathrm{e}=3.37\times 10^7`$ K, see Table 1). It is interesting to note that the IR flux density of the transient IR emission from the dust trail is predicted to rival that of the photon-heated dust in the galactic disk, and, despite the difference in heating mechanism, have similar colours (with a spectral peak in the 100-200$`\mu m`$ range). This, combined with removal of dust from the disk, and hence reduced internal extinction, will create a discrete system with brighter apparent blue magnitudes and a boosted spatially integrated IR flux density. If seen in a distant cluster, where the intracluster IR component could not be resolved from the disk component, this could create the illusion of a galaxy with an enhanced star-formation activity, even though the star-formation in the galaxy may actually be somewhat suppressed by the gas removal in reality. Since spirals are more spread over the whole cluster area, following the density-morphological relation (Dressler 1980), they are by far less numerous in the centre of the Virgo cluster. From the number density profiles of late-type galaxies in the Virgo Cluster and assuming a King profile with a core radius $`\mathrm{r}_\mathrm{c}=3.2^{}`$ (Binggeli et al. 1987) we estimate only a few spirals ($`2`$ galaxies) in the inner $`1^{}`$ from the cluster centre. As ram-pressure stripping is likely to be a transient phenomenon, occurring when the galaxies enter the dense core region of the IC medium, with sputtering timescales shorter than the time to traverse the angular extent of the core, we expect only of order 1 short-lived intracluster IR source in the inner $`1^{}`$ region of the cluster. To conclude, the IR emission coming from dust stripped from late-type galaxies is localised and connected to the parent galaxy, and does not account for a diffuse intracluster IR component. ## 4 An infall model for the dust We have shown that the diffuse infrared emission from dust produced by discrete sources inside the inner region of the Virgo cluster is below any detection capabilities. This situation occurs from the fact that there are simply not sufficient strong sources of dust to produce a dust injection rate high enough to replenish the large amount of dust sputtered by the ambient hot plasma. In this section we propose a new mechanism for injecting dust into the ICM, namely dust removed from spiral galaxies by galactic winds throughout their life time, and brought into the cluster by the infalling spirals. This scenario is a corollary of the fact that the clusters are still in the process of forming, by accreting galaxies from the surrounding field environment. The Virgo Cluster is a typical example (and best studied) for such a process. Tully & Shaya (1984) studied the phase space distribution of the spirals in and near the Virgo cluster and developed a mass distribution model in which galaxies within about 8 Mpc of the Virgo cluster are now falling back towards the cluster. Within their model most of the spirals entered the cluster in the last one-third to one-half of the age of the universe. This implies that the Virgo spirals were formed in lower density environments, more like field galaxies and have only lately entered the high-density cluster environment. Following the idea of clusters accreting from the field, we propose the following scenario (a schematic picture of this infall model is given in Fig. 5.): 1. Spiral galaxies are systematically infalling into the gravitational potential well of the cluster. Following Tully & Shaya (1984) we consider that all the spirals seen today in the Virgo Cluster have entered the cluster in the last 4-6 Gyr at a constant rate. Since the distribution of galaxies in the field is very clumpy, the infalling rate depends on the details of this distribution. But as we do not know the exact distribution of field galaxies over the last 4-6 Gyr, we can only assume a constant rate over this time. Furthermore, if spirals are destroyed or transformed into ellipticals when they approach the cluster core, then there may have been even more spirals infalling into the cluster, than we see today. Thus our assumptions give only a lower limit to the total number of infalling spirals. 2. In addition to the infalling galaxies, there should, presumably, be an accompanying primordial infalling diffuse intergalactic gas component, from which the galaxies originally formed. Little is known about the density and temperature of this intergalactic medium (IGM) in general, but it is believed to have temperatures in the range 10$`{}_{}{}^{4}10^6`$ K. Immediately, prior to galaxy formation one would expect the gas to co-move with the galaxies. At later stages of the infall, gas-gas hydrodynamic interactions may however decouple the gas from the galaxy inflow, and, provided the interactions are gentle enough to avoid heating to virial temperature, would tend to increase the infall rate by removing angular momentum. 3. The diffuse low pressure ambient IGM is favourable to the formation of galactic winds in spirals (Breitschwerdt et al. 1991). If we assume that the galactic winds of spirals were able to inject dust into the IG medium for their full lifetime of 13 Gyr, then large amount of dust is also brought into the cluster. A basic premise of our analysis is that the spirals infalling into the cluster will be approximately comoving with the (primarily primordial) IG gas and the injected dust and gas from the embedded galaxies. This premise can only be indirectly tested as any infalling IG medium can only be detected once it has been virialised into an X-ray emitting IC medium. Simple analytic models for the growth of clusters are given by Gunn & Gott (1972), in which the infalling baryonic material is relatively cold, such that the hot X-ray emitting IC medium is separated from the infalling gas by a strong accretion shock. In the present work, we tentatively identify the outer boundary of the faint diffuse X-ray emission emission extending $`45^{}`$ from M87 with a macroscopic accretion shock. The relatively sharp radial cut off of the diffuse X-ray emission shown particularly by the GINGA scan data (Takano et al. 1989) would seem to admit such an interpretation. This allows us to estimate the current upstream density of the infalling IG medium from the observed gas density of $`\mathrm{\hspace{0.17em}4}\times 10^5\mathrm{cm}^3`$ (Schindler et al. 1999; see also Sect. 2) for the boundary of the diffuse X-ray emitting medium. Taking the compression ratio of the accretion shock to be 4, this yields n$`{}_{\mathrm{IGM}}{}^{}\mathrm{\hspace{0.17em}10}^5\mathrm{cm}^3`$ for the current number density of the infalling IG medium at a distance of 1.3 Mpc from M87. We can now use this estimate for n<sub>IGM</sub> in conjunction with the observed total mass of baryonic matter emitting X-rays downstream of the accretion shock, which should be made up of infallen matter and the initial perturbation, to test the hypothesis that the infalling galaxies and IG gas are comoving. We can estimate radial infall velocity of the IG gas at the radius of the presumed accretion shock of R$`{}_{\mathrm{shock}}{}^{}1.27`$ Mpc (4) from the radial distribution of galaxian velocities derived by Tully & Shaya (1984) in their simple radial infall model (see their Fig. 4). After taking into account the slight differences in the assumed distance to the cluster (16.8 Mpc was derived by Tully & Shaya) we predict the gas infall velocity at the accretion shock to be v$`{}_{\mathrm{infall}}{}^{}\mathrm{\hspace{0.17em}800}`$km s<sup>-1</sup>. From this, we can estimate the current accretion rate of baryonic IG gas, assuming spherical symmetry, as $`4\pi `$R$`{}_{}{}^{2}{}_{\mathrm{shock}}{}^{}`$m$`_\mathrm{H}`$v$`_{\mathrm{infall}}`$n$`{}_{\mathrm{IGM}}{}^{}`$ $`3900`$M$`_{}`$yr<sup>-1</sup>. Since both the IC medium and the infalling galaxies show marked deviations from spherical symmetry, indicative of the infall of an irregular distribution of clumps, this estimate should be treated as a crude upper limit. Nevertheless, it is remarkable that the 3900 M$`_{}`$yr<sup>-1</sup> multiplied by a presumed age of $`1.3\times 10^{10}`$yr, or $`5.1\times 10^{13}`$M is close to the $`4\times 10^{13}`$M for the total baryonic gas mass in the (dominant) IC medium component centered on M87 within $`1.3`$Mpc, as plotted on Fig. 11a of Schindler et al. (1999). (Although Schindler et al. use a hydrostatic model for the derivation of mass from the X-ray emission, we would not expect very different results from a dynamic accretion flow model as the post shock inwards flow velocities are small). This result we use it here to justify our assumption of a radial inflow of relatively cold IG gas, comoving with the galaxies, which provides a medium to sweep grains injected from the embedded spiral galaxies down into the cluster. We also note that the incoming mass flux of the IG medium dominates that of the incoming galaxies, so that it is a safe assumption that the galactic winds themselves should have no significant heating effect on the IG medium. Here we consider the fate of dust particles ejected in the winds of infalling spiral galaxies. In the following we assume that grains always comove with the ambient gas - i.e. that the grains are charged and the ambient gas is magnetised. There are two main obstacles for such grains reaching the benign environment of the IGM, which is too cool for significant sputtering. Firstly, the grains have to survive the passage through the hot galactic wind; we will show in the next section that in most cases grains can survive this passage. Secondly, the winds will create local “bubbles” in the ambient IGM, containing wind shocks which provide a further possibility for sputtering. However, the densities are so small that this effect is negligible. According to Castor et al. (1975), the density and temperature in the hot shocked wind region is given by: $`\mathrm{n}_\mathrm{w}=0.01\mathrm{n}_0^{19/35}(\dot{\mathrm{M}}_6\mathrm{v}_{2000}^2)^{6/35}\mathrm{t}_6^{22/35}\mathrm{cm}^3`$ (20) $`\mathrm{T}_\mathrm{w}=1.6\times 10^6\mathrm{n}_0^{2/35}(\dot{\mathrm{M}}_6v_{2000}^2)^{8/35}\mathrm{t}_6^{6/35}\mathrm{K}`$ (21) where $`\dot{\mathrm{M}}_6=\dot{\mathrm{M}}_\mathrm{w}/(10^6\mathrm{M}_{}\mathrm{yr}^1)`$, $`\mathrm{v}_{2000}=\mathrm{v}_\mathrm{w}/(2000\mathrm{km}/\mathrm{s})`$, $`\mathrm{t}_6=\mathrm{t}_{\mathrm{gal}}/(10^6\mathrm{yr})`$, M<sub>w</sub> is the mass ejected through the winds during the age of the galaxy t<sub>gal</sub>, at a constant rate $`\dot{\mathrm{M}}_\mathrm{w}`$, v<sub>w</sub> is the terminal velocity of the wind, and n<sub>0</sub> is the number density of the ambient IGM. For typical values of $`\dot{\mathrm{M}}_\mathrm{w}=1\mathrm{M}_{}\mathrm{yr}^1`$, $`\mathrm{v}_\mathrm{w}=500`$ km/s (Breitschwerdt et al. 1991) and $`\mathrm{n}_0=10^5\mathrm{cm}^3`$ we obtain $`\mathrm{n}_\mathrm{w}=4\times 10^7\mathrm{cm}^3`$ and $`\mathrm{T}_\mathrm{w}=2.1\times 10^6`$K. For such a low density the sputtering time of a grain of size 0.1 $`\mu `$m is $`2.5\times 10^{11}`$ yr and therefore this effect can be neglected. A more fundamental effect is that the bubbles are buoyant, and will tend to rise relative to the cold ambient medium. This effect can potentially impede the inflow of the grains. However, even if the bubbles could grow for a substantial fraction of the Hubble time, it seems very unlikely that the velocity of a boyant bubble relative to the ambient medium could become significant compared to the inflow velocities. A fundamental upper limit for this relative velocity is the sound velocity of the ambient medium, which for T$`10^5`$ K is $`33\mathrm{km}\mathrm{s}^1`$. This is much smaller than the expected escape velocity for a central mass of $`2\times 10^{14}\mathrm{M}_{}`$ (Schindler et al. 1999), for the scales we consider here (out to 8 Mpc). In particular we note that the Virgocentric infall of the Local Group is 220 km/s (Tammann & Sandage 1985). In practice, as the galaxies approach the inner region of the cluster, they will decouple from the velocity field of the IGM medium. Then it seems unlikely that the bubbles will remain intact, due to the expected relative motions of the galaxies with respect to the primordial IGM. Analogous to stars moving through the ISM of the Milky Way, the bubbles would be expected to develop axi-symmetric “cometary” like structures, in which the hot material can escape into the ambient medium along the trailing axis. We can use the estimate of the density of the baryonic IG medium to check our assumption that sputtering can be neglected for grains released directly into the infalling IGM. Very conservatively, assuming that density varied with redshift as $`(1+`$z$`)^3`$ (i.e. ignoring the comoving density increase due to the infall into the cluster), n$`{}_{\mathrm{IGM}}{}^{}10^5\mathrm{cm}^3`$ corresponds to $`2.7\times 10^4`$ at z$`=2`$, for which (using Eq. (19) and taking the gas temperature to be constant in time at $`10^5`$ K), the sputtering timescale would be less than a Hubble time only for very small grains (of size $`\mathrm{\hspace{0.17em}0.002}\mu m`$). Thus it seems reasonable to suppose that grains can accumulate in the inflowing IG medium since rather early epochs. In summary, we would expect grains ejected in galactic winds to fall into the cluster, co-moving with the ambient intergalactic gas. In the cold ambient gas outside the X-ray emitting region of the cluster the grains will survive, since there is no mechanism for destruction. But once the gas and grains reach the X-ray region, the grains will be sputtered and collisionally heated, and thus produce infrared emission. In order to calculate the infrared emission from the intracluster medium, we must first consider destruction processes of grains in the galactic winds, which in principal could reduce the grain injection rate. ## 5 Grain sputtering and size distribution in the galactic winds From the study of the parameters of different kind of winds (Breitschwerdt et al. 1991, Breitschwerdt & Schmutzler 1999) we know that the initial conditions in the winds are rather unfavourable to grain survival, due to the high densities and temperatures. Thus the grain survival depends on how rapidly these winds can cool and expand on their way out of the disk. There are mainly two types of galactic winds in the spiral galaxies: winds driven by cosmic rays and thermally driven winds (star-burst winds, like in M82). Cosmic ray winds can be classified (Breitschwerdt & Schmutzler 1999) in global winds, that originate from the large-scale expansion of the hot galactic corona, and local winds, which come from individual superbubble regions in the disk. Star-burst winds are stronger, being thermally driven due to the high gas temperature in their hot interstellar medium. Here we will consider the parameters of these different kinds of winds, as presented in Breitschwerdt & Schmutzler (1999), who included in their calculations both adiabatic and radiative cooling. Especially for global winds it was shown that the radiative cooling dominates the gas cooling until the temperature drops to a few $`10^4`$ K, when adiabatic cooling becomes important. Again, considering global winds Zirakashvili et al. (1996) have shown that unsaturated non-linear Landau damping may dominate the advection of waves in the plasma and hence lead to local dissipative heating. Up to now there is no self consistent model for galactic winds that includes both the effects of line cooling and heating due to wave damping. Below we argue that the contribution of wave heating is much smaller than that of cooling, and can be neglected. Thus, we have calculated the cooling rate and the heating rate using the solutions for global winds from Breitschwerdt & Schmutzler (1999). The cooling rate C is: $`\mathrm{C}=\mathrm{n}^2\mathrm{\Lambda }(\mathrm{T})`$ (22) where n is the gas density in cm<sup>-3</sup> and the cooling function $`\mathrm{\Lambda }(\mathrm{T})`$ was taken from Kahn (1976), $`\mathrm{\Lambda }=1.33\times 10^{19}\mathrm{T}^{1/2}(\mathrm{erg}\mathrm{cm}^3\mathrm{s}^1)`$ (23) The upper limit to the heating rate H is: $`\mathrm{H}=\mathrm{v}_\mathrm{a}\mathrm{P}_\mathrm{c}`$ (24) with v<sub>a</sub> the Alfvén velocity and P<sub>c</sub> the cosmic ray pressure. The effects of heating and cooling can be compared in Fig. 6. Above 1 kpc from the disk, the cooling rate is 3 orders of magnitudes higher, and it is still dominant at 10 kpc. At even higher distances the contribution of the two rates becomes equal, but the gas temperature and density is already below the level where it could produce any efficient dust grain destruction. Therefore we will consider in our calculation the wind solution from Breitschwerdt & Schmutzler (1999). ### 5.1 Global winds In the slow global winds that are continuously emitted by the quiescent spiral galaxies, the temperature drops very quickly with the expansion of the wind, and this reduces substantially the grain sputtering. The temperature (T \[K\]), density (n \[cm$`{}_{}{}^{3}]`$), and velocity (v \[km/s\]) profiles above the disk (z\[kpc\]) are plotted with solid line in Fig. 7, using the data from Breitschwerdt & Schmutzler (1999) (also from D. Breitschwerdt, private communication). The global wind solutions correspond to a flux tube located at a galactocentric distance of 10 kpc, for an adiabatic model whereas an isochoric cooling function was included. These profiles were used for calculating the sputtering time profile, considering the analytical formula from Tsai & Mathews (1995). Thus, the rate at which the radius of the dust grain decreases with time in a hot plasma of temperature T and density n<sub>H</sub> is: $`{\displaystyle \frac{\mathrm{da}}{\mathrm{dt}}}=f\mathrm{n}_\mathrm{H}\left[\left({\displaystyle \frac{\mathrm{T}_\mathrm{d}}{\mathrm{T}}}\right)^{2.5}+1\right]^1`$ (25) Tsai & Mathews showed that this relation is a good approximation to the detailed calculations of Draine & Salpeter (1979) and Tielens et al. (1994), for both graphite and silicate when $`f=3.2\times 10^{18}\mathrm{cm}^4\mathrm{s}^1`$, and T$`{}_{\mathrm{d}}{}^{}=2\times 10^6`$ K. The sputtering time can be then computed as: $`\mathrm{t}_{\mathrm{sput}}=\mathrm{a}\left|{\displaystyle \frac{\mathrm{da}}{\mathrm{dt}}}\right|^1`$ (26) which for gas temperatures $`10^6<\mathrm{T}<10^9`$ K reduces to the formula (5). In Fig. 7 we show the sputtering time for a big grain of radius $`\mathrm{a}=0.25\mu `$m. t<sub>sput</sub> increases rapidly to more than $`10^{10}`$ yr, but smaller grains have shorter survival times. The minimum grain size that can survive in the wind is derived from the cumulative size loss $`\mathrm{\Delta }`$a, which is obtained by integrating the sputtering rate over the time it takes the wind to go out the galaxy: $`\mathrm{\Delta }\mathrm{a}={\displaystyle \frac{\mathrm{da}}{\mathrm{dt}}\mathrm{dt}}`$ (27) The cumulative size loss $`\mathrm{\Delta }`$a is plotted in Fig. 8 (solid line), and from the saturation value we derive a minimum grain size that survives sputtering $`\mathrm{a}_{\mathrm{surv}}=0.0071\mu `$m. We assume that the initial size distribution in the wind is given again by a power law of index k=3.5. The sputtering modifies the size distribution, in the sense that grains smaller than a<sub>surv</sub> will be completely destroyed while bigger grains will end with a final size $`\mathrm{a}\mathrm{a}_{\mathrm{surv}}`$. The final size distribution will be given by: $`\mathrm{N}(\mathrm{a})(\mathrm{a}+\mathrm{a}_{\mathrm{surv}})^k`$ (28) In the case of global winds grains as small as $`0.0072\mu `$m can survive, which means that the grain size distribution is only negligible changed, and most of the dust will be injected in the IC medium. ### 5.2 Local winds Local winds have initially higher densities and temperatures, and these quantities remain almost constant till the wind reaches 10 kpc above the galactic disk (Breitschwerdt & Schmutzler 1999). The corresponding temperature, density and velocity profiles are plotted in Fig. 7 with short-dashed lines. We expect that less grains will survive the wind. Following the same recipe like in the case of global winds we obtain a minimum grain size that survives, $`\mathrm{a}_{\mathrm{surv}}=0.084\mu `$m. This is higher than the minimum grain size that survives the global winds, but still a substantial amount of grains will survive. The corresponding sputtering time and cumulative size loss $`\mathrm{\Delta }`$a are plotted with short-dashed lines in Fig. 7 and 8, respectively. ### 5.3 Star-burst winds The very energetic starburst winds can be traced already from 0.1 kpc above the disk, but in Fig. 7 and 8 we plot (long-dashed lines) only the range between 1 and 100 kpc, in order to have the same range as for the other two type of winds. The temperature and density is more than one order of magnitude higher than for global winds, and even at 10 kpc these parameters are still high enough to sputter quite efficiently the dust grains. Despite the fact that the wind velocity is also very high, the minimum grain size that survives is $`\mathrm{\Delta }\mathrm{a}=0.3\mu `$m. But such big grains are rare in the diffuse interstellar medium. We conclude that star-burst winds are not able to inject dust into the IC medium. ## 6 The emission spectrum produced by dust falling into the cluster core We have seen that dust grains survive mainly in global winds, and that their size distribution is not significantly changed by sputtering in the injection process. If the initial dust-to-gas ratio in the wind is Z$`{}_{\mathrm{d}}{}^{}=0.0075`$ (the Galactic value), the grains will be injected into the IG medium with Z$`{}_{\mathrm{d}}{}^{}=0.0071`$ after sputtering, which means that the dust-to-gas ratio is practically unchanged. The mass loss rate through global galactic winds is of the order of $`1\mathrm{M}_{}/\mathrm{yr}`$ for a galaxy like the Milky Way (Breitschwerdt et al. 1991, Zirakashvili et al. 1996). The local winds can also inject grains into the IC medium, but with a smaller dust-to gas ratio, and with a modified size distribution given by (22). Furthermore, the mass loss rate from these winds is one order of magnitude lower than for the global winds (Breitschwerdt & Schmutzler 1999). And local winds do not happen continuously over the life time of the spirals, but rather in episodes that are connected with the life time of their parent giant HII region superbubbles. Thus we neglect their contribution. Star-burst winds can inject large amounts of gas in the IC medium, even more than 10 M/yr, but the injected gas is depleted of grains. To conclude, spiral galaxies release grains in the IC medium mainly through global galactic winds. ### 6.1 Injection rate of grains into the intracluster medium We estimate the total infalling rate of grains into the cluster by scaling the mass loss rate in galactic winds to the blue luminosity of infalling spirals. Following Tully & Shaya (1984) (see Sect. 4) let us suppose that all galaxies currently seen within 6 of M87 arrived at a constant rate within the last 5 Gyr.<sup>1</sup><sup>1</sup>1We note that this may actually yield a lower rate than the actual rate if some fraction of the infalling spirals had been transformed into ellipticals in this time. These observed spirals have a total blue luminosity of $`7.15\times 10^{11}\mathrm{L}_{}`$ (Binggeli et al. 1987), so that we can express the infalling rate of luminosity as 1.43$`\times 10^{11}\mathrm{L}_{}\mathrm{Gyr}^1`$. We further assume that the galactic winds were able to inject dust into the IG medium for their full lifetime of 13 Gyr. Then, scaling to the wind injection rate of 1 M$`_{}`$yr<sup>-1</sup> for the Milky Way (L$`{}_{\mathrm{B}}{}^{}\mathrm{\hspace{0.17em}1.6}\times 10^{10}\mathrm{L}_{}`$; de Vaucouleurs & Pence 1978) as estimated by Breitschwerdt et al. (1991), each 1.6$`\times 10^{10}\mathrm{L}_{}`$ entering the cluster will bring with it an accumulated wind-ejected gas mass of 13$`\times 10^9\mathrm{M}_{}`$. These estimates then lead to a gas infall rate from galactic winds of 116 M$`_{}`$yr<sup>-1</sup>. To calculate the corresponding grain infall rate we take the dust-to-gas mass ratio in the ejected winds, averaged over the 13 Gyr, to be the solar value of 0.0075 to yield a dust injection rate of 0.87 M$`_{}`$yr<sup>-1</sup>. This may be quite conservative, bearing in mind that the abundances of the galactic winds should be strongly enhanced compared to the ambient interstellar medium (Völk 1991). In addition to dust injected into the external IGM by galaxies infalling for the first time, there may also be a secondary contribution from spirals which entered the hot IC medium at previous epochs, but have orbits taking them back beyond the accretion shock, where they can again inject dust.<sup>2</sup><sup>2</sup>2We assume winds are suppressed while the galaxies are in the pressurised IC medium. The maximum time since the currently seen spirals entered the cluster ($``$5 Gyr) is not much longer than the $``$3.2 Gyr needed to traverse the volume enclose by the accretion shock on a radial path. Thus, only some fraction - perhaps one third - of the spirals will have passed back through the accretion shock. Their residence times after re-entry depends on the actual orbits and are thus very uncertain, but may be of the order of 2 Gyr, in any case much less than the $``$13 Gyr for the freshly infalling objects. If one third of the galaxies contribute grains to the inflow for 2 Gyr after reentry, the gas and dust injection rates calculated above (for the freshly infalling systems) should be augmented by $``$5 percent to 122 and 0.91 M$`_{}`$yr<sup>-1</sup> respectively. Finally, to calculate the dust injection rate into the IC medium, we should take account of the fact that the accretion shock is not stationary, but advancing outwards in the cluster reference frame at some low speed v<sub>adv</sub>. Thus the true injection rate of grains into the hot IC medium will be somewhat greater than that given by the infall rate (as calculated above) by a factor $`\zeta =`$ (v$`{}_{\mathrm{adv}}{}^{}+`$v<sub>infall</sub>)/v<sub>infall</sub>. A rigorous calculation of $`\zeta `$ is beyond the scope of this paper, but an upper limit - for the case that the accretion shock had a compression ratio of 4 and the material downstream of the accretion shock was almost stationary in the cluster reference frame - would be 1.33.<sup>3</sup><sup>3</sup>3In reality the downstream flow clearly has to have some inwards velocity to maintain pressure equilibrium against gravity in the IC medium. Another estimate for $`\zeta `$ can be obtained from the estimate of v$`{}_{\mathrm{infall}}{}^{}800`$ km s<sup>-1</sup> from the infall of the galaxies (following Tully & Shaya (1984) - see Sect. 4) and a crude value for v<sub>adv</sub> of $`800`$ km s<sup>-1</sup> which is the accretion shock radius (1.27 Mpc) divided by the age of the cluster (13 Gyr). This yields $`\zeta `$1.125. We will assume $`\zeta =`$1.1, from which we obtain final estimates of 134 and 1.0 M$`_{}`$yr<sup>-1</sup> for the gas and dust injection rates into the IC medium. ### 6.2 Heating and sputtering of injected grains In order to calculate the heating of grains injected from the inflowing IG medium we recall consider the X-ray morphology of the Virgo Cluster, as summarised in Sect. 2. We only consider the dominant M87 subcluster, which accounts for 71$`\%`$ of the total emission out to 4-5. This consists of the dense hot X-ray core region extending 1 around M87, and the faint diffuse emission extending 4-5 from M87, which produces 15$`\%`$ of the total X-ray flux (Böhringer et al. 1994). We have identified the boundary of the faint diffuse emission as an accretion shock (Sect. 4). Grains injected at the supposed accretion shock should be heated by a plasma of density $`\mathrm{\hspace{0.17em}4}\times 10^5\mathrm{cm}^3`$ and temperature $`3\times 10^7`$K. The sputtering timescales in this most tenuous region of the IC medium are 2.5$`\times 10^9`$ yr of grains of size 0.1 $`\mu m`$. Although comparable to the infall time for galaxies from the accretion shock radius to the centre, this sputtering timescale does not mean that the IR emission will be distributed over a broad range of radii. The grains will follow the inflow of the gas downstream of the accretion shock, which should be actually quasi hydrostatic. Thus, in this simple homogeneous picture for the cluster the grains would be expected to be heated and sputtered in the vicinity of the accretion shock 4-5 degrees from the cluster core, and the infrared emission would trace the morphology of the accretion shock surface. In this scenario, the grain heating will be relatively low as the grains never reach the denser core regions of the IC medium, and the emission will predominantly arise in the submillimetre range. We refer to this as case A. In reality the structure of the M87 subcluster is much more irregular than this simple representation. Böhringer et al. (1994) have shown that in the western part of the cluster the X-ray emission falls off more steeply, while the northern edge of the cluster is less well defined, with the cluster boundary dissolving into several individual subclumps. Because of the uncertainties in the 3D structure of the cluster, it is very difficult to predict the actual fate of infalling grains reaching the accretion shock radius. In particular, the accretion may be fundamentally clumpy in nature, with infalling clumps of dust-bearing gas reaching different depths before interacting and merging with an irregularly shaped intracluster medium. In this scenario it may be possible for clumps infalling through certain position angles to directly interact with the dense X-ray core region extended 1 around M87. Then the grain heating will be stronger, with most radiation being produced in the FIR. We refer to this as case B. We note that the instantaneous luminosity will be very similar which ever of case A and B is nearer the truth, provided the grain injection rate is constant over the sputtering timescale. For a steady state between injection and destruction, the effect of the increased heating expected for grains reaching the central dense regions of the IC medium (Case B) will almost exactly be balanced by the shorter grain survival times, since sputtering timescales are only weakly dependent on temperature for plasma hotter than $``$10$`^6`$K. The spectrum of the submillimeter emission in case A is given in Fig. 9, both for silicate grains (dashed line) and for graphite grains (dashed-dotted line). In the calculation we assume a steady-state solution for the balance between dust destruction and injection just downstream of the accretion shock, following the procedure described in Sect. 2. The peak of the emission, the colour of the spectrum and the fluxes are given in Table 4. Fig. 9 also shows the predicted emission from case B (solid line for silicates and dotted line for graphites). Here the spectrum was calculated under the assumption that the grains will survive and emit only in the outer spherical shell of the central core of cluster at radii of $``$1 (n$`{}_{\mathrm{e}}{}^{}=0.0007\mathrm{cm}^3`$, T$`{}_{\mathrm{e}}{}^{}=3.37\times 10^7`$, see Table 1). The parameters of this emission are also given in Table 4. ## 7 Discussion We have calculated the infrared emission for the Virgo cluster taking into account all possible sources of dust inside the cluster core. We have shown that there are not enough discrete sources of dust in the Virgo cluster to produce detectable diffuse emission. This is compounded by the fact that galaxian sources of dust embedded in the IC medium will certainly not be able to provide a smooth distribution of grains within the volume of the IC medium as sputtering timescales are invariably shorter than transport timescales over a typical separation between galaxies. Of all potential discrete sources only the stars could be thought of as giving rise to a smooth emission component, though, for the Virgo cluster, this will only amount to 20 percent of the emission from the infalling intergalactic grains. Thus, any detection of truly diffuse FIR emission is likely not to trace dust injected by galaxies inside the cluster core, but rather the inflow of grains to the cluster from the external intergalactic medium. Since younger clusters, like the Virgo cluster still have spiral galaxies infalling into the cluster, we might expect them to have a larger amount of infalling dust, and thus a higher FIR emission. Thus, in general, the phenomenon of diffuse intracluster FIR emission may give information on the current dynamic age of the cluster. In this sense it is complementary to the measurements of X-ray emission which broadly relate to all the baryonic matter accumulated over the lifetime of the cluster. ### 7.1 Detectability and recognition of IC dust emission Our estimates for the IR emission rely on a chain of quite poorly determined quantities, such as the dependence of mass ejection in galactic winds on blue luminosity, uncertainties in the dynamical properties of the inflow, and the metallicity of the wind ejected material. Detection of diffuse IR emission may be useful to constrain some of these factors. On the basis of the simple estimates given in Table 4, however, it will be fundamentally difficult to detect such emission from nearby clusters. The high angular sizes will lead to severe confusion of the IC IR emission with foreground emission, which will be difficult to alleviate with multiwavelength studies given that the predicted emission has quite similar colours to the galactic cirrus. The best possibility for detection is for our case B scenario in which the cluster is fundamentally clumpy, as then the emission may be concentrated over smaller solid angles (scales of 1 as opposed to 4 for the case A scenario). In general, because the grain sputtering timescales are shorter than the grain transport timescales, once the grains have reached their sputtering/injection sites, they will in general trace surfaces of injection rather then be distributed over a depth of volume. This may lead to some limb-brightening effects which could aid detection, though care would have to be taken to distinguish these structures from the trails predicted from spiral galaxies undergoing ram-pressure stripping discussed in Sect. 3.3. It is more likely that diffuse intracluster emission could be detected towards distant clusters of lower angular size. Referring again to our illustrative scenarios of grains being sputtered in the diffuse X-ray emitting plasma (Case A) or when they first enter the 1 degree core region of the cluster (Case B), we address the question of how the Virgo cluster would appear if situated at a cosmological distance of z=0.5. The 4 degree region of the cluster would be seen as a source of 3. So the cluster will look like a compact source, and to detect this source we need to have only enough total flux. If we consider a Euclidian flat geometry with q<sub>0</sub>=1/2 we obtain a flux of 1.5 mJy (redshifted at $`\lambda =389\mu `$m) for Case A or a flux of 1 mJy (redshifted at $`\lambda =263\mu `$m) for Case B. There will be of course the emission from dust within the cluster galaxies. Here, the main confusion problem would be with the FIR emission from constituent spiral galaxies. To estimate the galaxy contribution to the total FIR emission we used the correlation of the B magnitude of 105 disk galaxies selected from the Virgo Cluster Catalogue with their IRAS 100 micron emission. We found S$`{}_{\nu }{}^{}(100\mu m)\mathrm{S}_\nu (\mathrm{B})^{1.25}`$ in mJy. Using again a King profile for the Virgo Cluster spiral galaxies we derived a total blue luminosity within the 4 degrees region from the cluster centre of L$`{}_{\mathrm{B}}{}^{}=3.7\times 10^{11}\mathrm{L}_{}`$. This translates into a flux of 295 Jy at 100 micron. For a colour F$`{}_{175}{}^{}/\mathrm{F}_{100}=1.5`$ we estimate that the cluster galaxies should contribute with 442 Jy to the total flux at 160 $`\mu `$m. This translates into a flux of 13 mJy for a redshifted cluster, which is still one order of magnitude higher than the intracluster emission. Thus, a combination of good surface brightness performance and resolution will be required for a unambiguous detection of IC dust emission. Especially as the emission peaks in the submillimeter, this type of observation will be particularly well suited to the new generation of submillimeter interferometers. Finally, we remark that galaxies may not be the only sources of IG grains, especially in the early universe. FIR/ sub-mm observations of clusters may in general provide new information of the abundance of the IG grains, which is currently only poorly constrained. ### 7.2 The optical extinction in clusters From our infalling model it is obvious that the extinction in clusters will be dominated by dust upstream the accretion shock. If we suppose that grains are uniformly distributed within the volume of the cluster, we obtain an A$`{}_{\mathrm{B}}{}^{}=0.005`$ magnitudes at a distance 4 degrees from the cluster center. However, this extinction could increase when some limb brightness effects are taken into account. Furthermore, extinction may be higher on some optical paths if we allow for clumpy accretion of infallen material. In passing we mention that these calculations are obtained under our very conservative assumption that galactic winds have the Galactic gas-to-dust ratio. As discussed in Sect. 6.1, the abundance of galactic winds may be strongly enhanced as compared to the ambient interstellar medium (Völk 1991) and thus much more dust can then be injected in the IG medium. Thus the amount of extinction may be sensitive to the assumed accretion rate of dust and to the geometrical effects in the cluster. Our predicted extinction should be compared to the few tenths of optical depth predicted by some optical studies which found a deficit of distant galaxies (Zwicky 1962; Karachentsev & Lipovetskii 1969; Bogart & Wagoner 1973; Szalay et al. 1989) or quasars (Boyle et al. 1988; Romani & Maoz 1992). However there are also some optical studies that do not support the presence of extinction in clusters. Maoz (1995) found out that radio-selected quasars behind Abell clusters are not redder than quasars that do not have a cluster in foreground. He also suggested that previous results that claimed the avoidance of foreground Abell clusters by optically selected quasars can be explained as being due to selection effects. Ferguson (1993) studied the colour versus Mg<sub>2</sub> relation for elliptical galaxies in cluster, group and field samples. He also found no evidence for excess reddening in clusters or groups. ### 7.3 Relation to other observables Our predictions for diffuse FIR/sub-mm intracluster emission rely on the ubiquitous existence of “global” winds from quiescent spiral galaxies, enriching the infalling intergalactic medium with metals, partly as dust grains. As noted by Völk (1991), it should be expected that the metallicity of the ejected wind material should be substantially higher than for the interstellar medium of the disks, which has consequences for the chemical evolution of spirals. An interesting corollary, therefore, of any detection of diffuse IR emission in the periphery of the cluster, is the implication for the environmental effects on the chemical evolution of spiral galaxies. There is some evidence (Skillman et al. 1996) that spiral galaxies in the core region of the Virgo cluster have enhanced metallicities compared to counterparts in the cluster periphery and field environment. Skillman et al. (1996) account for this abundance contrast in terms of scenarios invoking the accretion of metal-poor cold intergalactic material to dilute the abundances of field galaxies, by supposing that the accretion is suppressed for galaxies in the cluster core due to the hot environment. We propose the opposite scenario to explain the abundance contrast between field and cluster, namely the ejection of overabundant interstellar material in quiescent winds of field galaxies, as opposed to the cluster core, where quiescent winds are predicted to be suppressed by the high pressure environment. We note that a wind scenario for chemical abundance regulation would require a larger integrated nucleosynthetic production over the lifetime of a galaxy as compared to a regulation by accretion, due to the loss of a substantial amount of metals to the IG medium. The infall scenario may also alleviate the problem of explaining the absolute metal content of the intracluster medium purely in terms of the nucleosynthetic history of galaxies, principally ellipticals (e.g. Okazaki et al. 1993) embedded in the IC medium, as some proportion of the metals will then have been injected through the winds of spirals embedded in the infalling IG medium. Recently, Wiebe et al. (1999) also considered the possibility that galactic winds from spiral galaxies contribute to the chemical enrichment of the ICM, though they did not take into account the fact that winds are suppressed in the dense ICM by the high pressure environment. In our infall model, one possible observational manifestation would be the expected X-ray emission in the periphery region of the cluster, just downstream of the accretion shock. Once sputtered, the grains will release refractory elements into the gas phase, which will radiate line emission to supplement emission from the directly injected non-refractory component. The radial profiles of line emission may give some indication of the infall history of these metals. Our predictions for the origins of diffuse infrared emission in the IC medium may also have a physical corollary to the observational situation regarding the diffuse synchrotron emission. Grains are short-lived and cannot move far from their sources. But by the same token, relativistic electron cannot move far from their sources due to inverse Compton and synchrotron losses. It is tempting to imagine a scenario in which both phenomena are localised around an accretion shock and would be interesting to compare the morphology of diffuse IR and radio emission from this viewpoint. The possibility of particle acceleration at the location of extended accretion shocks around the clusters has been recently discussed by Enßlin et al. (1998). They identified the so called cluster radio relics to be powered by the accretion shock produced by the large-scale gas motion around the clusters. Finally, we can examine whether submillimeter emission from intracluster dust, as predicted here, might be detectable in experiments to detect Sunyaev-Zeldovich (SZ) excess towards clusters. At increasing redshifts, the wavelength of peak emission from the IC dust may move close to the SZ peak around $`\lambda =0.8`$ mm. The appearance of both the intracluster dust emission and the SZ excess is predicted to be related to the morphology of the diffuse X-ray Bremsstrahlung from the ICM. To get an estimate for the relative emissions from the two phenomena we have estimated the (thermal) SZ effect that we would observe for a cluster like the Virgo Cluster, using the parameters for the 4 degree central region of Virgo, but with the cluster redshifted to z=0.5. The Comptonization parameter y (see Rephaeli 1995) was calculated by integrating the gas density profile of the cluster (Schindler et al. 1999) for a constant temperature $`\mathrm{T}=3.3\times 10^7`$ K. We obtained an averaged $`\mathrm{y}=4.3\times 10^6`$, which is consistent with the upper limit determined from the COBE/FIRAS database, $`\mathrm{y}2.5\times 10^5`$ (Mather et al. 1994). The corresponding temperature change due to the scattering was calculated in the non-relativistic case (Rephaeli 1995): $`\mathrm{\Delta }T_{\mathrm{nr}}=\left[{\displaystyle \frac{x(e^x+1)}{e^x1}}4\right]\mathrm{T}_0\mathrm{y}`$ (29) where x=h$`\nu `$/kT is the nondimensional frequency, T is the radiation temperature and $`\mathrm{T}_0=2.726`$ (as determined by the COBE/FIRAS; Mather et al. 1994). The corresponding surface brightness due to this change in temperature was multiplied with a solid angle of $`3^{}`$ radius (as seen for the 4 degrees inner region of the Virgo Cluster at a redshift z=0.5) and the integrated flux is shown in Fig. 10. The fluxes estimated for the diffuse dust emission (in both Case A and Case B) for the Virgo Cluster (redshifted to z=0.5) are also plotted for comparison in Fig. 10. It is obvious that for wavelengths less than 400 micron the diffuse emission is dominated by dust emission. At longer wavelengths the dust emission becomes less important, reaching about 5 percent of the S-Z emission at the S-Z peak. However, these considerations relate to the integrated fluxes. If the diffuse dust emission could be spatially resolved, as indeed it must be to distinguish it from emission from discrete sources in the cluster, it may be possible to distinguish it from the SZ emission due to it’s predicted limb brightening at the boundary of the accretion shock. Indeed, in the outermost regions of the cluster, IC dust emission may give a significant signal compared to the smoothly, centrally peaked SZ emission. ## 8 Summary In this work we show that any detection of a diffuse FIR intracluster emission is likely not to trace dust injected by galaxies inside the cluster core, since there are not enough discrete sources of dust to produce detectable emission. We propose that the diffuse intracluster FIR emission may trace the current accretion rate of the cluster, which give information on the current dynamic age of the cluster. Even in this case the estimated amount of dust in the cluster is lower than suggested in earlier work (Dwek et al. 1990). The main results of this paper are listed below: * The infrared emission from dust injected by intergalactic stars inside the core region of the Virgo cluster is a factor of $`10`$ below the detection limit of currently available telescopes. * The amount of intracluster infrared emission produced by dust ejected by early type galaxies in the inner core of Virgo cluster is negligible even in the upper limit calculation. * The IR emission coming from dust stripped from late-type galaxies is localised and connected to the parent galaxy, and does not account for a diffuse intracluster IR component. * For dynamically young clusters like the Virgo cluster we propose a new mechanism for injecting dust in the ICM, namely dust removed from spiral galaxies by galactic winds throughout their life time, and brought into the cluster by the IGM. A basic premise of this scenario is that the spirals infalling into the cluster comove with their ambient IG medium and the injected dust and gas from the embedded spirals. * In this work we identify the outer boundary of the diffuse X-ray emission of the Virgo Cluster, extending $`45^{}`$ from M87, with a macroscopic accretion shock. In a simple homogeneous picture for the cluster the infalling grains would trace the morphology of the accretion shock surface. Since the density of the ambient plasma is low in this region, the heating is relatively low and the infrared emission arise in the submillimetre range. If the accretion is fundamentally clumpy in nature, it is possible for clumps infalling through certain position angles to directly interact with the dense X-ray core region of the cluster. Then the infrared emission is shifted to shorter wavelengths. * For nearby clusters it will be fundamentally difficult to detect infrared emission from intracluster dust, due the large volumes considered and to the relative low masses of dust. It is more likely that diffuse intracluster emission could be detected for distant clusters. Because the cluster emission is dominated by emission from cluster member galaxies, any detection of a diffuse component requires a combination of good surface brightness performance and resolution. This type of observation would be well suited to the new generation of submillimeter interferometers. ###### Acknowledgements. We would like to thank Dr. M. Voit for the careful refereeing of the manuscript. We gratefully acknowledge Dr. D. Breitschwerdt for providing us with his calculated data on galactic winds. We also kindly acknowledge Dr. U. Fritze-von Alvensleben for providing us with the HRD population model needed in Sect. 3.1. We would like to thank Drs. A. Burkert and A. Blanchard for interesting discussions.
warning/0001/nucl-th0001015.html
ar5iv
text
# Light-particle emission from the fissioning nuclei 126Ba, 188Pt and 266,272,278110: theoretical predictions and experimental results The work is partly supported by the Polish Committee of Scientific Research under Contract No. 2P03B 011 12 ## 1 Introduction The dynamical time evolution of the fission process from an initially formed compound nucleus with a more or less compact shape to the saddle and scission configurations and the simultaneous emission during this deformation process of light particles constitutes a very complex problem. It is the aim of our study to describe this process taking into account the excitation of both thermal and rotational nature, the initial configuration and to calculate the evaporation of light particles from an excited, rotating, deformed nucleus on its way from the initial compact shape to the scission point. In the absence of a complete microscopic ab initio theory of such a dynamical process, many different theoretical approaches aiming at its description have been worked out -. They generally rely on a classical description of the evolution of the collective coordinates which are introduced as abundant variables. In these descriptions, collective parameters appear (collective mass, friction, and diffusion coefficients) which depend on the collective coordinates. It seems clear that the quality of the theoretical description will depend on the more or less pertinent choice of the collective coordinates and the degree of realism of the underlying theory used to determine the collective parameter functions. The importance of pure quantum effects such as pairing correlations and the existence of shell effects which are present at low excitation energies has been investigated recently -. The multiplicity and the characteristics (energies, angular distributions, correlation functions) of prefission light particles and gamma emission may provide some useful information on the time evolution of the nucleus as it evolves towards the saddle and scission point. Observables like prefission light-particle multiplicities may, indeed, work as a clock. Angular distributions and particle correlations may give informations on the surface and deformation of the fissioning nucleus. We recently developed a realistic time dependent model which described the time evolution of an excited nucleus as it is e.g. generated in a heavy-ion collision and decaying through symmetric fission with pre- and post-fission light particle emission . We applied our description to the case of <sup>160</sup>Yb and made a detailed analysis of different physical quantities characterizing the decaying system. We found a rather good agreement of our theoretical predictions with the experimentally observed light-particle multiplicities. More recently, several new sets of experimental data have become available - with a more complete analysis of the emitted light particles. It is the aim of the present work to systematically test our approach by a comparison with these experimental data. The fact that these data are obtained in quite different regions of the periodic table makes a comparison between calculated and experimental results all the more challenging. A particular motivation for this work is the explicit description of the entrance channel through which the initial spin distribution of the system is generated. This is achieved by considering all possible relative orientations of the two colliding deformed nuclei in the entrance channel. Their different orientations and the different possible impact parameters of the reaction yield the initial spin distribution $`d\sigma /dL`$ versus $`L`$ which determines the relative weights of the initial angular momentum of the compound nucleus with which the Langevin trajectories are started. Still another motivation for the present work is the presentation of the link between the well known Weisskopf formula and a new and more microscopically founded description of the emission width for the light particles in terms of phase-space densities of the particle to be emitted . The paper is organized as follows: The dynamical model describing the time evolution of a deformed, hot, and rotating nucleus from its initial to a final configuration is presented in section 2. In section 3 we specify how to determine the emission rates for light particles from a hot, deformed and rotating nucleus moving along the previously described trajectory. In section 4 we show how we obtain the initial spin distribution which has been not considered in our previous study . Our results concerning the nuclear systems <sup>126</sup>Ba, <sup>188</sup>Pt and <sup>266,272,278</sup>110 are presented in section 5. The paper is closed in section 6 with a summary of the most relevant results and an outlook on planned extensions of the model. ## 2 Collective dynamics of an excited system We consider an ensemble of deformed nuclei with finite excitation energies and rotational angular momenta as given by the initial conditions determined from the entrance channel. The subsequent time evolution of the nucleus is governed in our present description by a single collective coordinate $`q=R_{12}/R_0`$ where $`R_{12}`$ is the distance between the two centers of mass of the left-right symmetric deformed nucleus and $`R_0`$ is the radius of the corresponding spherical nucleus having the same volume. This collective variable is defined in the framework of a Trentalange–Koonin–Sierk (TKS) parameterization of the surface of the nucleus. The TKS deformation parameters are related to $`q`$ by means of a minimization procedure of the collective potential energy defined below . Denoting the conjugate momentum by $`p(t)`$ we use the following classical equations of motion to describe the time evolution of the fissioning nucleus $`{\displaystyle \frac{dq}{dt}}={\displaystyle \frac{p}{M(q)}}`$ (1) $`{\displaystyle \frac{dp}{dt}}={\displaystyle \frac{1}{2}}\left({\displaystyle \frac{p}{M(q)}}\right)^2{\displaystyle \frac{dM(q)}{dq}}{\displaystyle \frac{dV(q)}{dq}}{\displaystyle \frac{\gamma (q)}{M(q)}}p+F_L(t).`$ (2) Here $`M(q)`$ is the collective mass determined in the incompressible fluid approximation and $`\gamma (q)`$ the friction coefficient calculated in the wall-and-window friction model . The collective potential V(q) is obtained as the difference of the Helmholtz free energy at deformation $`q`$ minus the one for the ground-state deformation. The free energy could in principle be obtained from a microscopic mean-field calculation at finite temperature. To perform such a constraint Hartree-Fock calculation using a reasonable effective nucleon-nucleon interaction of the Skyrme or Gogny type at every point in the multidimensional deformation space is, of course, completely out of question due to the tremendous computer time such an analysis would involve. Even to perform the same kind of calculation on the level of a selfconsistent semiclassical approximation like the Extended Thomas-Fermi (ETF) method at finite temperature , which would describe the average nuclear structure without shell oscillations, would be far too time consuming. This is why we have rather used a still simpler semiclassical approach and the liquid drop model with the parametrization of Myers and Swiatecki as explained in Appendix A. We have tested the barrier heights obtained in this way versus the ones obtained from selfconsistent semiclassical calculations (at zero temperature) and obtained agreement within a few MeV as described in Ref.. These calculations describe, of course, deformation properties of non excited nuclei. In order to take nuclear excitation into account we have used the temperature dependence of the LDM parameters associated with the Skyrme SkM interaction , determined for that interaction through selfconsistent semiclassical calculations at finite temperature in Ref. . To use the temperature dependence of the LDM parameters associated with the Skyrme SkM force together with the Myers-Swiatecki parametrization of the LDM is, of course, inconsistent. Since approaches give very similar results for the semiclassical energy at zero temperature this approximation seems, however, very reasonable. The semiclassical approach used here is of course only an approximation at low temperatures since shell effects are absent from our description. It is however well known that nuclear shell effects are washed out with increasing nuclear temperature and have essentially disappeared beyond T$``$ 2.5-3 MeV. At these temperatures the semiclasssical results are becoming exact. For the cold systems the fission barriers obtained with this potential are higher that those evaluated by Sierk , but already at excitation energies $`E^{}90`$ MeV both barriers become comparable and for $`E^{}>90`$ MeV the barriers evaluated in our model are even smaller than the Yukawa folded ones of Sierk. The friction term and the Langevin force $`F_L(t)`$ in Eq. (2) generate the irreversible production of heat energy and the energy fluctuations respectively which both originate from the coupling of the collective dynamics to the intrinsic degrees of freedom. In practice one defines $`F_L(t)=\sqrt{D(q)}f_l(t)`$, where $`D(q)`$ is the diffusion coefficient. We take the simplifying point of view that it is related to the friction coefficient $`\gamma (q)`$ through the Einstein relation $`D(q)=\gamma (q)T`$, where $`T`$ is the temperature of the system. The quantity $`f_L(\tau )`$ can be written in the form $`f_L(\tau )=\sqrt{\tau }\eta `$, where $`\tau `$ is a time step length corresponding to a time interval $`[t,t+\tau ]`$ and $`\eta `$ is a gaussian distributed random number with zero average $`\eta =0`$ and variance $`\eta ^2=2`$ where brackets represent ensemble averages. The Einstein relation is in principle valid at high temperature, in a regime where the process can be described in a classical framework. Quantum effects (pairing, shell corrections, collective shape vibrations) may be present at low temperatures and modify the relation between $`D`$ and $`\gamma `$ -. We shall come back to this point later on. In principle, one would have to treat each ensemble of nuclei with a given initial angular momentum $`L`$ and a given initial excitation energy microcanonically. We assume that we may replace this microcanonical ensemble by a grand canonical one in which, instead of the mean excitation energy, the temperature $`T`$ has a well-defined value at a given time. This simplifying assumption is innocuous as far as the mean values of observables are concerned, but it may falsify the fluctuations to a certain extent. We assume that the time scale which governs the fission dynamics is much larger than the internal equilibration time, otherwise the definition of a nuclear temperature would make no sense. Under these conditions the system can be considered as being continuously close to equilibrium. We suppose that the nuclear excitation energy $`E^{}`$ is related to $`T`$ through the usual Fermi-gas expressions $`E^{}=a(q)T^2`$, where $`a(q)`$ is the level density parameter at a given deformation $`q`$. For practical calculations Eqs. (1) and (2) are rewritten in a discretized form and numerically integrated over small time steps $`\tau `$ . In order to perform such an integration one needs to start from some fixed initial conditions which define the beginning of the process. We take into account the possible emission of light particles in every time step and we make sure that the average total energy of the system will be conserved. This determines the nuclear excitation energy and hence the temperature at each instant of time. ## 3 Particle emission from a hot, deformed, and rotating nucleus Particle emission before scission is governed by transition rates $`\mathrm{\Gamma }_\nu ^{\alpha \beta }(E^{},L)`$ which determine the number of particles of type $`\nu `$ (we take into account neutrons, protons, and $`\alpha `$ particles) emitted per unit time with an energy $`e_\alpha `$ in an interval $`[e_\alpha \frac{1}{2}\mathrm{\Delta }e,e_\alpha +\frac{1}{2}\mathrm{\Delta }e]`$ and with an angular momentum $`\mathrm{}_\beta `$ from a nucleus with average excitation energy $`E^{}`$ and total angular momentum $`L`$. In Ref. we used the well known Weisskopf formula for the partial width $`\mathrm{\Gamma }_\nu ^{\alpha \beta }(E^{},L)`$ in terms of densities of states of the emitting and residual nucleus and of the transmission coefficient $`w_\nu (e,\mathrm{}_\beta )`$ for emitted a particle $`\nu `$ with given energy $`e`$ and angular momentum $`\mathrm{}_\beta `$. The determination of $`w_\nu `$ takes into account the deformation and rotation of the emitting nucleus (see Appendix B of Ref. for details). In Ref. another more microscopic determination of these transition rates was proposed. In this framework the transition rates $`\mathrm{\Gamma }_\nu ^{\alpha \beta }`$ are given as $$\mathrm{\Gamma }_\nu ^{\alpha \beta }=\frac{d^2n_\nu }{d\epsilon _\alpha d\mathrm{}_\beta }\mathrm{\Delta }\epsilon \mathrm{\Delta }\mathrm{},$$ (3) where $`\epsilon _\alpha `$ and $`\mathrm{}_\beta `$ characterize an emission energy and angular momentum lying in the intervals $$[\epsilon _\alpha \frac{1}{2}\mathrm{\Delta }\epsilon ,\epsilon _\alpha +\frac{1}{2}\mathrm{\Delta }\epsilon ]\mathrm{and}[\mathrm{}_\beta \frac{1}{2}\mathrm{\Delta }\mathrm{},\mathrm{}_\beta +\frac{1}{2}\mathrm{\Delta }\mathrm{}]$$ respectively. The number $`n_\nu `$ of particles of type $`\nu `$ which are emitted per time unit through the surface $`\mathrm{\Sigma }`$ of the fissioning nucleus is given by $$n_\nu =\underset{\mathrm{\Sigma }}{}𝑑\sigma d^3p^{}f_\nu (\stackrel{}{r}_0^{},\stackrel{}{p}^{})v_{}^{}(\stackrel{}{r}_0^{})w_\nu (v_{}^{}(\stackrel{}{r}_0^{}))$$ (4) where $`\stackrel{}{p}^{}`$, $`\stackrel{}{v}^{}`$ are the momentum and velocity in the body-fixed frame. The quantity $`v_{}^{}`$ is the velocity component perpendicular to the emission surface at the surface point $`\stackrel{}{r}_0^{}`$. The $`\stackrel{}{p}^{}=m\stackrel{}{v}^{}+m\stackrel{}{\omega }\times \stackrel{}{r}^{}`$ is the momentum of the particle of mass $`m`$ in the laboratory reference frame and $`\stackrel{}{\omega }`$ the angular velocity of the nucleus in this frame. Here and henceforward, primed quantities refer to the body-fixed frame. The classical distribution in phase space reads $$f_\nu (\stackrel{}{r}^{},\stackrel{}{p}^{})=\frac{2}{h^3}\frac{\theta (\stackrel{}{r}^{})}{1+\mathrm{exp}\left[\frac{1}{T}\left(\frac{p^2}{2m}+U\omega \mathrm{}^{}\mu _\nu \right)\right]}.$$ (5) The $`\theta `$ function is 1 if $`\stackrel{}{r}^{}`$ lies inside the nuclear volume or on its surface $`\mathrm{\Sigma }`$ and zero otherwise. The quantity $`\mu _\nu `$ is the chemical potential and $`\mathrm{}^{}`$ the body-fixed angular momentum in the direction perpendicular to the axis of rotational symmetry of the deformed nucleus. The potential $`U`$ is taken as $$U(\stackrel{}{r}^{})=V_0+V_{Cb}(\stackrel{}{r}^{}),$$ (6) where $`V_0>0`$ is chosen as a constant mean field potential and $`V_{\mathrm{Cb}}`$ is the Coulomb potential experienced at $`\stackrel{}{r}^{}`$ by protons. The quantity $`w_\nu (v_{}^{}(\stackrel{}{r}_0^{}))`$ is the classical transmission coefficient for the emission of a particle of type $`\nu `$. The transmission factor $`w_\nu `$ was chosen to be the one of an inverted harmonic oscillator . The explicit relation of (3) with the Weisskopf formulation is discussed in Appendix B for a spherical emitting nucleus with angular momentum $`L=0`$. The rates given by (3) can be worked out numerically if the distribution function in the phase space is known, what is the case for the neutrons and protons but not for the $`\alpha `$-particles. We are working on a model which describes the distribution of $`\alpha `$-particles by folding of the distributions of four (2$`n`$ and 2$`p`$) correlated particles. The present approach is like the one of Ref. and it allows us, in principle, to determine the particle emission in a given direction of space, hence the determination of observables like angular distributions and particle-particle correlation functions. Such observables may be worked out in the future. We did not attempt to do it here because of the lack of corresponding experimental data. The transition rates $`\mathrm{\Gamma }_\nu ^{\alpha \beta }`$ are used in a simulation algorithm by means of which we determine at each time step $`[t,t+\tau ]`$ along each classical trajectory whether a particle of given type with an energy and angular momentum in given intervals is emitted or not from the compound nucleus. Since the algorithm is already described in detail in Ref. we do not repeat it here. ### 3.1 Initial conditions and energy balance In order to integrate the classical equations (1) and (2) we need to fix the initial conditions from which the compound system starts and evolves either through a fission channel to its saddle and scission point or stays as a compound system which only emits light particles, i.e. ends up as an evaporation residue. All the experimental systems which are considered in the next section are generated by means of heavy ion collisions at some bombarding energy. The nuclei which are involved in the reaction process can be deformed. The initial conditions corresponding to the origin of time are fixed by $`q_0`$ and $`p_0`$, the initial value of the collective variable and its conjugate momentum and the spin distribution of the system which fixes the relative weight of the angular momentum of the initial compound systems. The coordinate $`q_0`$ is fixed at the value of $`q`$ where the collective potential $`V(q)`$ is minimal and its conjugate momentum is drawn from a normalized gaussian distribution $$P(p_0)=(2\pi MT_0)^{1/2}\mathrm{exp}(p_0^2/2MT_0),$$ (7) where $`M=M(q_0)`$ is the collective mass. The initial temperature $`T_0`$ is obtained through $`E_0^{}=a(q_0)T_0^2`$, where $`E_0^{}`$ is the initial excitation energy which can be obtained from the knowledge of the total energy as explained below. In the reaction process, the compound nucleus can be formed with different values of the angular momentum. If both nuclei are spherical it is easy to construct the initial spin distribution under the assumption that the reaction cross section is given by the geometrical expression $$\frac{d\sigma (L)}{dL}=\frac{\pi }{k^2}(2L+1),$$ (8) where $`k`$ is the wavenumber of the relative motion. The situation is more complicated if one or both initial nuclei are deformed. In principle one should then consider all possible relative orientations of the nuclei and follow their relative trajectories from an infinite distance up to fusion. This implies the knowledge of the nuclear and Coulomb potential for given relative distance between the centers of mass and given orientation. This raises intricate problems which have been considered by several authors but not yet solved in their full generality. In a first step we want to avoid such cumbersome calculations and restrict ourselves to a simplified procedure which has been already proposed in former work . In practice we consider the scattering of spherical ions of fixed energy $`E`$ and introduce the gaussian energy distribution with the width $$\mathrm{\Delta }E=\delta E+\mathrm{\Delta }_1E+\mathrm{\Delta }_2E,$$ (9) where $`\delta E`$ is the experimental beam width, $`\mathrm{\Delta }_1E`$ describes the slowing down of the projectile in the target, and the quantity $`\mathrm{\Delta }_2E`$ is the difference of the Coulomb barriers between two extreme geometries of touching spheroidally deformed nuclei, i.e. with aligned (tip on tip) or parallel (side on side) $`z`$–axes. One can now work out the classical trajectories describing the relative motion of the equivalent spherical ions for every initial angular momentum $`L`$ in $`[0,L_{\mathrm{max}}]`$, where $`L_{\mathrm{max}}`$ is the maximal angular momentum for the colliding ions and for a sufficient number of energy values within the distribution defined above. Repeating the trajectory calculations leads to the spin distribution $$\left(\frac{d\sigma _F}{dL}\right)_{L_i}=\frac{2\pi }{k^2}L_i\frac{N_i^F}{N_i},$$ (10) where $`L_i`$ is the considered angular momentum, $`N_i^F`$ is the number of trajectories which lead to fusion and $`N_i`$ is the total number of trajectories. The quantity $`k`$ is the wave–number of the relative motion of the incident nuclei. The present procedure, even though it is not rigorous, leads to spin distributions which are rather realistic . Finally, in order to describe the evolution of the excited compound system one needs to follow the evolution of the average excitation energy along the trajectory. This is achieved by the requirement of energy conservation. At the initial point of a trajectory the total available energy can be written as $`E_{\mathrm{tot}}`$ $`=`$ $`E_{\mathrm{coll}}+E_{\mathrm{rot}}+E_0^{}`$ (11) $`=`$ $`{\displaystyle \frac{p_0^2}{2M(q_0)}}+V(q_0)+{\displaystyle \frac{L_0^2}{2J(q_0)}}+E_0^{}.`$ Here, $`J(q_0)`$ is the moment of inertia of the compound system taken as a rigid deformed rotator and $`E_0^{}`$ is the initial intrinsic excitation energy of the compound nucleus. The intrinsic excitation energy $`E^{}(t)`$ at any given later time $`t`$ can be determined from the energy balance $$E_{\mathrm{tot}}=E_{\mathrm{coll}}+E_{\mathrm{rot}}+E_0^{}+B_\nu +e_\alpha +E_{\mathrm{recoil}},$$ (12) where $`B_\nu `$ is the binding energy of the emitted particle, different from zero only for $`\alpha `$-particles, $`e_\alpha `$ is the kinetic energy of the emitted particle and $`E_{\mathrm{recoil}}`$ is the recoil energy of the nucleus after the emission of a particle, which can be neglected in practical calculations. For each choice of the initial conditions one generates a separate trajectory. Emitted particles are counted with their energy and angular momentum. If the system overcomes the fission barrier the trajectory (event) contributes to the final fission cross section $$\sigma _{\mathrm{fiss}}=\underset{i}{}\frac{d\sigma _{\mathrm{fiss}}}{dL_i}=\underset{i}{}\left(\frac{d\sigma _F}{dL_i}\right)_{L_i}\frac{N_i^{\mathrm{fiss}}}{N_i^F},$$ (13) where $`N_i^{\mathrm{fiss}}`$ is the number of trajectories which lead to fission, and $`N_i^F`$ – the number of fused trajectories at a given angular momentum $`L_i`$. The sum runs over all angular momentum bins and $`d\sigma _F/dL`$ is the fusion cross section given by Eq. (10). The numbers of prefission particles obtained for each angular momentum of the compound nucleus ($`M_\nu `$) are weighed with the differential fission cross section in order to obtain the measured number of particles emitted in coincidence with fission: $$<M_\nu >=\frac{\underset{i}{}\frac{d\sigma _{\mathrm{fiss}}}{dL_i}M_\nu (L_i)}{\sigma _{\mathrm{fiss}}}.$$ (14) ## 4 Application of the model to various experimentally studied systems In the present section we aim to confront experimental results concerning the multiplicities of emitted particles in coincidence with symmetric fission events with the model described in the preceding section. All measured and calculated data are presented in Table 1 for three systems: <sup>126</sup>Ba, <sup>188</sup>Pt, and <sup>266,272,278</sup>110 obtained by different entrance channels and at different energies. These data result from experiments performed at the SARA (Grenoble) and the VIVITRON (Strasbourg) - using the DEMON neutron detector . Unfortunately, the charged particles ($`p,\alpha `$) were not measured, so we give in Table 1 the theoretical predictions only. We present a more detailed discussion of the results for each system in the next subsections. ### 4.1 The <sup>126</sup>Ba compound system The compound nucleus <sup>126</sup>Ba has been experimentally produced through two different entrance channels: (I) $`{}_{}{}^{28}\mathrm{Si}+^{98}\mathrm{Mo}^{126}\mathrm{Ba}`$ at $`E_{lab}`$=204, 187, and 166 MeV and (II) $`{}_{}{}^{19}\mathrm{F}+^{107}\mathrm{Ag}^{126}\mathrm{Ba}`$ at $`E_{lab}`$=148 and 128 MeV. Fig. 1 shows the fusion cross-section and the fission yields as a function of the initial angular momentum. The distributions of the initial angular momentum were calculated with the help of the corresponding Langevin equation for two different entrance channels (I, II) and various bombarding energies. As it can be seen, the fission yields are rather small and located in the tails of the distributions. This observation which also holds for many other systems implies that it is of great importance to calculate carefully the dependence of the fusion cross-section on the angular momentum if one wants to describe the competition between the decay of the compound nucleus by fission and by light particle emission in a correct way. Concerning Fig. 1 we also note that the dominant part of the initial excitation energy resides in the collective rotational motion. One observes in Fig. 1 a general increase of the fission cross section with the excitation energy of the compound nucleus <sup>126</sup>Ba with the only exception of the highest excitation energy. In that case the emission of $`\alpha `$-particles becomes important and competes with the fission process. This causes a substantial loss of internal energy due to particle emission which in turn leads to an increase of the fission barrier and therefore to a smaller fission yield. In Fig. 2, the potential energy $`V_{\mathrm{fiss}}`$ of the fissioning nucleus <sup>126</sup>Ba for different angular momenta is plotted as a function of the relative distance between fission fragments. One can see in Fig. 2 that the fission barrier of the fused nucleus at temperature $`T`$=1.6 MeV and angular $`L`$=0 is quite high in our model. It is equal to 43 MeV. It vanishes at very large angular momentum ($`L80\mathrm{}`$). The high temperature can also reduce significantly the fission barrier as one can learn from Fig. 3, where the free energy of <sup>126</sup>Ba is plotted for a few temperatures between 0 and 4.8 MeV. This is why one can observe a significant fission rate only at large angular momenta and high excitation energies. All curves representing the fission barriers in Figs. 2 and 3 end at the scission point region. It is seen from Figs. 2 and 3 that the saddle and scission points are close to each other for <sup>126</sup>Ba. Part of the excitation energy of <sup>126</sup>Ba is contained in the rotational mode. In Fig. 4, the temperature of <sup>126</sup>Ba corresponding to different excitation energies is plotted as a function of angular momentum. One can see that, at the lowest excitation energy $`E^{}=`$84 MeV and the highest angular momentum, the temperature is 0.5 MeV only. This means that our theoretical predictions based on the statistical equilibrium could be too rough in this case. The experimental neutron multiplicities obtained for both the fission and the fusion channel are shown in Table 1 along with the calculated neutron, proton, and $`\alpha `$-particle multiplicities. In Figs. 5 and 6, we show the calculated average numbers of neutrons, protons, and $`\alpha `$-particles emitted in coincidence with fission as a function of the excitation energy $`E^{}`$ of the initial compound nucleus. The initial compound nucleus <sup>126</sup>Ba is produced by the fusion reaction (I) in Fig. 5, and in Fig. 6 the same initial compound nucleus originates from the fusion reaction (II). Only the $`n`$-emission on the way to fission has been measured so far. The measured numbers of prefission neutrons are indicated by points with error bars. At two excitation energies we have measurements for the two different entrance channels. As one can read from Table 1 at the excitation energy $`E^{}`$ = 101.4 MeV the measured and calculated numbers of emitted prefission neutrons (in coincidence with fission) are $`M_n`$ = 1.32 and 1.83, resp., for the entrance channel (I) (Fig. 5) and $`M_n`$ = 1.31 and 1.80, resp., for the entrance channel II (Fig. 6). At the higher excitation energy $`E^{}`$ = 118.5 MeV the measured and calculated numbers of emitted neutrons are $`M_n=2.01`$ and 1.71 for the entrance channel (I) (Fig. 5) and $`M_n=1.85`$ and 1.99 for the entrance channel (II) (Fig. 6). The experimentally observed number of emitted prefission neutrons at given excitation energy $`E^{}`$ of the initial compound nucleus is thus larger for the entrance channel (I) than for the entrance channel (II) and the difference of the number of emitted neutrons is seen to be larger for the larger excitation energy. The interpretation of this observation is that for <sup>126</sup>Ba only the compound nuclei formed with high angular momentum may undergo fission and thus give rise to prefission neutrons because the fission barrier for low angular momentum is high ($``$ 15 MeV). For $`T`$=1.6 MeV the LD fission barriers vanish for about $`L=70\mathrm{}`$ as can be seen from Fig. 2. Thus the main part of the prefission neutrons is obtained for angular momenta around $`L=80\mathrm{}`$ as indicated in Fig. 1. At given excitation energy $`E^{}`$ the number of compound nuclei formed with such high angular momentum is larger for the entrance channel (I) with <sup>28</sup>Si as a projectile than for the entrance channel (II) with <sup>19</sup>F as a projectile (see Fig. 1). The theoretically estimated prefission proton multiplicities ($`M_p`$) are very small as one can see in Figs. 5 and 6. The multiplicities of $`\alpha `$-particles are not negligible at the highest ($`E^{}`$=131.7 MeV) and the lowest ($`E^{}`$=84.1 MeV) excitation energies (see Figs. 5 and 6). The increase of $`M_\alpha `$ at high excitation energy is mostly due to the high temperature effects while the unexpectedly large number of emitted $`\alpha `$-particles at $`E^{}`$=84.1 MeV is mostly due to the large deformation of the emitter and the huge centrifugal potential ($`L80\mathrm{}`$) reducing the already small Coulomb barrier for the $`\alpha `$-particles at the tips even further. As for the comparison of measured and calculated numbers of emitted prefission neutrons, the general agreement is not unsatisfactory. Nevertheless, one notices that the calculated average emission numbers are larger than the observed ones for the higher excitation energy $`E^{}`$ = 118,5 MeV. If, instead of the temperature-independent friction, which underlies these calculations, we would use a friction parameter increasing with temperature, the agreement might be better. ### 4.2 The <sup>188</sup>Pt compound system As seen in Table 1, the compound nucleus <sup>188</sup>Pt was produced at two excitation energies $`E^{}`$=100 MeV and 66.5 MeV using two different reactions * $`{}_{}{}^{34}\mathrm{S}+{}_{}{}^{154}\mathrm{Sm}^{188}\mathrm{Pt}`$ at $`E_{lab}`$= 203 and 160 MeV , * $`{}_{}{}^{16}\mathrm{O}+{}_{}{}^{172}\mathrm{Yb}^{188}\mathrm{Pt}`$ at $`E_{lab}`$= 138 MeV. The theoretical estimates of the initial spin distributions (solid lines) as well as the fission rates (bars) for $`E^{}`$=100 MeV and two entrance channels as well as for $`E^{}`$=66.5 MeV are plotted in Fig. 7 as functions of angular momentum of the compound nucleus. It is seen that each reaction leads to different spin distribution. These differences are responsible for the entrance channel effects in the prefission neutron multiplicities emitted by <sup>188</sup>Pt at $`E^{}`$=100 MeV (see Table 1). Similarly, as in the case of <sup>126</sup>Ba, we present in Fig. 8 the fission barriers of <sup>188</sup>Pt obtained for different angular momenta ($`L`$) at fixed temperature $`T`$=1.6 MeV. It is seen that for $`T=1.6`$ MeV and $`L=70\mathrm{}`$ the fission barrier becomes negligible. Also one can see in Fig. 9 that with increasing temperature the fission barrier decreases. At $`T=4.8`$ MeV and $`L=30\mathrm{}`$, the fission barrier of <sup>188</sup>Pt is about four times lower than for $`T=0`$ and $`L=30\mathrm{}`$. Contrary to the case of <sup>126</sup>Ba, where the saddle point is very close to the scission point, the path from the saddle point ($`q=R_{12}/R_0`$ 1.6) to the scission point ($`q`$ 2.2) of <sup>188</sup>Pt is much longer and will take more time. This implies that the case of <sup>188</sup>Pt is better suited to study the influence of dynamical effects on the prefission light-particle multiplicities. The prescission neutron multiplicities theoretically predicted for <sup>188</sup>Pt are too small (by $``$1 unit on the average) in comparison with the experimental data . Also the effect of the entrance channel is not fully reproduced. Unfortunately the protons and $`\alpha `$-particle multiplicities were not measured, so we cannot test the predictive power of our model in this case. ### 4.3 The $`Z=110`$, $`N=156,162,`$ and $`168`$ compound systems Three isotopes of the superheavy compound nucleus with $`Z=110`$ were formed by means of the following fusion reactions * <sup>58</sup>Ni + <sup>208</sup>Pb $``$ <sup>266</sup>110 * <sup>64</sup>Ni + <sup>208</sup>Pb $``$ <sup>272</sup>110 * <sup>40</sup>Ca + <sup>232</sup>Th $``$ <sup>272</sup>110 * <sup>40</sup>Ar + <sup>238</sup>U $``$ <sup>278</sup>110 at different excitation energies between 66 MeV and 186 MeV. The prefission neutron multiplicities were determined experimentally and estimated theoretically within our model. Both the experimental and theoretical results are given in Table 1. Using Langevin type equations to describe the fusion process we obtained the initial spin distribution of compound nucleus. As an example, the differential fusion cross section is plotted in Fig. 10 for the reaction <sup>40</sup>Ca+ <sup>232</sup>Th at $`E_{lab}`$=250.4 MeV as a function of the angular momentum of the system. The effective fission cross section is marked by bars in the figure. It is seen in Fig. 10 that only the lowest angular momenta contribute to the fusion fission process of <sup>272</sup>110 due to the fact that the fission barrier of <sup>272</sup>110 vanishes at $`L>22\mathrm{}`$. All isotopes of the element Z=110 have a rather small fission barrier which disappears rapidly with increasing spin ($`L`$). This is illustrated in Fig. 11, where the fission barriers of <sup>266</sup>110 corresponding to a few $`L`$ are plotted as functions of the elongation. In our calculations we only took into account those trajectories for which the saddle point existed, i.e. we neglected the so–called quasifission events. The number of emitted prescission particles is mostly governed by the dynamics from the saddle to the scission point, but of course the initial conditions (see Section 3.1) play also an important role. In our model we used two different sets of Langevin equations. The first set of equations describes the fusion dynamics while the second one, coupled with the Master equations for particle emission, describes the fission process . A better solution would be a model in which the whole dynamics from fusion to fission as well as the light particles emission would be described by one set of equations. Such an approach could immediately solve the problem of the proper setting of the initial conditions for the compound nucleus. The experimental and theoretical prefission neutron multiplicities for $`Z=`$110 compound nuclei are written in Table 1 and additionally they are drawn in Fig. 12 as functions of the excitation energy of the system. The dashed lines correspond to the results obtained with the wall-and-window friction ($`\gamma _{ww}`$) reduced by 50% while the dotted lines stand for the results obtained with the standard value of $`\gamma _{ww}`$. It is seen that for the lowest excitation energies the agreement of theoretical results with the measurements of neutron multiplicities is good, while the discrepancy grows with increasing excitation energy. For the most excited compound nuclei our theoretical predictions are too large by 1 to 2 units. This result could mean that the effective time which the system needs to pass from the saddle to the scission point is too long in our model for a very hot compound nucleus. Any of transport coefficients do not depend on the temperature. This could be the reason of the observed discrepancy. In the near future we intend to perform a new calculation with temperature dependent transport coefficients evaluated within the linear response theory . The temperature dependence of the friction parameter could be important, because as one can see in Fig. 12 the reduction of the friction parameter by 50% decrease the neutron multiplicities by 1 to 0.5 units depending on the excitation energy of compound nucleus. In our model the superheavy nucleus with Z=110 is already cold when it reaches the scission point after emission of some neutrons and $`\alpha `$-particles. So the shell effects begin to play an important role for very deformed shapes. The shell effects could lead to the compact fission, i.e. splitting of the Z=110 nucleus into two spherical fragments. In our model the shell effects are not present but we simulate partially their influence on the number prefission particles by counting only those particles which have been evaporated before reaching the elongation of nucleus close to the compact scission point ($`R_{12}/R_01.5`$). Contrary to the case of decay of <sup>126</sup>Ba and <sup>188</sup>Pt compound nuclei a hot superheavy nucleus with Z=110 exists and fissions at very low angular momenta ($`L<20\mathrm{}`$). This range of L corresponds to the linear part of the differential fusion cross-section $`d\sigma /dL`$ (see Fig. 10), so we do not expect entrance channel effects on multiplicities of prefission neutrons. The Coulomb barrier for emission of the charged particles from $`Z=`$110 isotopes is rather high so the predicted proton and $`\alpha `$ multiplicities are much smaller than the neutron multiplicities as one can see in Table 1. ## 5 Summary, conclusions and further developments In the first part of the present work we described an extension of our model used so far for describing the fission dynamics of an excited, rotating and possibly deformed compound system. We considered the decay of compound nuclei produced by the fusion of two heavy ions. The initial spin distribution was determined by a simple model calculation which took the possible deformation of the two ions into account . Furthermore, we introduced the microscopic classical expression for the emission rates of light particles derived in reference and showed the link of this expression with the Weisskopf formulation . In the second part of the paper we presented a detailed comparison of the calculated neutron multiplicities with the ones obtained in different experiments involving systems from different regions of the mass table. All five systems discussed in the paper (<sup>126</sup>Ba, <sup>188</sup>Pt and Z=110 isotopes) are good examples of different fission and particle evaporation mechanisms. The light nucleus <sup>126</sup>Ba fissions at very high angular momenta ($`L80\mathrm{}`$). Its saddle point is close to the scission point, so the fission dynamics plays a rather minor role in this case. The entrance channel effects influence significantly the decay properties of the excited <sup>126</sup>Ba, via different initial spin distributions of compound nucleus. Also huge centrifugal forces and big deformations corresponding to the saddle point lead in this case to a true competition between the neutron and $`\alpha `$-particle evaporation. The experimental prefission neutron multiplicities grow with the excitation energy of <sup>126</sup>Ba while it is not always the case in our theoretical estimates. In the case of <sup>188</sup>Pt the average angular momentum of the fissioning nucleus is around $`60\mathrm{}`$, i.e. $`20\mathrm{}`$ less than for <sup>126</sup>Ba. Also the way from the saddle point to the scission point is longer so that the dynamics from the saddle to the scission plays here an important role. In addition, the evaporation of a few particles does not increase so dramatically the height of the fission barrier as in the case for <sup>126</sup>Ba. As a consequence the fission process takes place during a longer time and the number of evaporated particles is larger. Similarly as in the case of <sup>126</sup>Ba the initial spin distribution (the entrance channel effect) influences the multiplicity of pre-fission neutrons which is here too small (by $``$ 30%) in comparison with the experimental data. A completely different case is the decay of superheavy compound nuclei with Z=110. Here the way from the weakly deformed saddle point to the scission point is very long. The fission barrier of a hot compound nucleus with Z=110 is very small and it vanishes already at low angular momenta ($`L<22\mathrm{}`$). The way in which the Z=110 nuclei has been produced does not influence their decay properties since such nuclei exist at small angular momenta only, which belong to the linear part of the fusion cross-section $`d\sigma /dL`$. A majority of compound nuclei goes to fission with simultaneous emission of neutrons and $`\alpha `$-particles. The number of prefission neutrons depends in this case on details of the fission dynamics, e.g. the slope of the potential and the magnitude of the friction and diffusion parameters. Our model overestimates slightly ($`15`$%) the experimental number of prefission neutrons here. The neutron emission process cools significantly the Z=110 compound nuclei, so the temperature dependence of the friction and diffusion parameters could be “visible” in this case. The present formulation of the model and its application can be improved on several points: * One may introduce a more detailed description of the formation process of the compound nuclei. Such a description will involve explicit trajectory calculations of the incident nuclei taking into account the nuclear and the Coulomb potentials for different relative initial orientations of the deformed nuclei. Calculations of relative potentials have been worked out in this framework but, to our knowledge, a full fledged dynamical trajectory calculation has yet to be done. * The full treatment of fusion fission dynamics with simultaneous emission of the light particles will release us from a somewhat arbitrary choice of the initial conditions for the compound nucleus. Such a project is at present under investigation. * We used a classical description of fluctuations and a form of the fluctuation-dissipation theorem which are valid for the case of high temperature. In fact, the temperature of the systems which we considered above could be low enough for quantum effects (pairing and shell effects) to play a non negligible role. This point has been investigated in the recent past within the framework of nuclear transport theory . It is our aim to come back to this point in the near future. * Our fission barrier estimates need also further improvements. The present model bases on a modified (see Appendix A) liquid drop formula of Myers and Swiatecki and it is known that for the cold light nuclei like <sup>126</sup>Ba it overestimates the fission barriers by $`10`$ MeV . For reasons of comparison with experiment we shall generalize our model taking asymmetric fission events into account. Many compound nuclei generated from mass asymmetric formation process channels appear with a high statistical weight. Last but not least it would be nice to extend the test of the model to observables other than the light particle emission rates. Out-of plane particle angular distributions and correlations between emitted light particles would provide further sensitive tests of the validity of our approach. For instance, the ratio of neutrons emitted out of the reaction plane and within the reaction plane (defined by the beam and the outgoing fission fragments) would enable us to determine the deformation of the emitting compound nucleus. Furthermore, we stress the importance of a measurement of the emitted neutrons, protons, and $`\alpha `$-particles in the same system because the emission rates of these particles are reciprocally dependent on each other. Acknowledgements The authors J. B., K. D., and J. R. wish to express their thanks for the warm hospitality extended to them at the Institute for Theoretical Physics of the UMCS of Lublin. Appendix A: Transport parameters used in the Langevin equation Fission dynamics correlated with prefission particle emission is generally described in a phenomenological framework, by means of quantities such as collective temperature-dependent potentials, masses, moments of inertia and friction coefficients. The fission process is described here in terms of the unique collective variable $`q=R_{12}/R_0`$. The fission path parameterized by $`q`$ is fixed by means of a minimalization of the free energy of a nucleus in the $`(\alpha _0,\alpha _2,\alpha _4,\alpha _6)`$ deformation space at a fixed temperature $`T`$. The deformation parameters are defined as follows $$\rho _s^2(z)=\mathrm{R}_0^2\underset{\mathrm{}=0}{\overset{6}{}}\alpha _2\mathrm{}(\mathrm{q})\mathrm{P}_2\mathrm{}(\frac{\mathrm{z}}{\mathrm{z}_0}),\mathrm{z}_0\mathrm{z}\mathrm{z}_0,$$ $`(\mathrm{A1})`$ where the function $`\rho _s(z)`$ is the distance in cylindrical coordinates of any point of the surface to the symmetry axis. The Helmholtz free energy of the nucleus plays the role of the collective potential and can be written, in a semiclassical approximation, as $$F(N,Z,q,L,T)=E(N,Z,q,T=0)a(N,Z,q)T^2+E_{rot}(N,Z,q,L).$$ $`(A2)`$ The first term on the r.h.s. of eq. (A2) is a liquid-drop (or any macroscopic model) type energy expressing the semiclassical energy of the deformed cold nuclear system as a function of the mass number $`A=N+Z,`$ and the isospin asymmetry $`I=(NZ)/A`$. The deformation dependence of the free energy is taken into account through the shape functions $`B_s`$ and $`B_{coul}`$. The rotational energy is calculated explicitly as $`E_{rot}=L^2/2𝒥(q)`$ with a $`q`$-dependent rigid-body moment of inertia. We have used in eq. (A2) the liquid drop parametrization of Myers and Swiatecki . The effects of excitation are taken into account in eq. (A2) through the level-density parameter $`a`$ at a given deformation. As the Helmholtz free energy, which is the variational quantity, is given by $$F(T)=E(T)TS(T)$$ $`(A3)`$ with the temperature T as a Langrange multiplier, one can show that for a liquid-drop type system, entropy and excitation energy read $$S=2aT$$ $`(A4)`$ and $$E^{}=E(T)E(T=0)=aT^2$$ $`(A5)`$ from which one obtains the last relation in eq. (A2). The level density parameter $`a`$ can be written in the form $$a(N,Z,q)=a_v(1+k_vI^2)A+a_s(1+k_sI^2)A^{2/3}B_s(q)+a_{coul}Z^2A^{1/3}B_{coul}(q),$$ $`(A6)`$ where the parameters $`a_v`$=0.0533 MeV<sup>-1</sup>, $`k_v`$=0.5261, $`a_s`$=0.1059 MeV<sup>-1</sup>, $`k_s`$=2.7192, and $`a_{Coul}`$=0.000458 MeV<sup>-1</sup> are taken from Ref. . In this expression, as well as in the macroscopic energy in eq. (A2), we have neglected the curvature and compression term proportional to $`A^{1/3}`$ since its coefficient for the cold system is known experimentally to be small . When evaluating the Coulomb term in (A2) and (A6) the Coulomb exchange contribution is neglected. To describe the fission dynamics one also needs to calculate the mass parameter $`M(q)`$ which enters the kinetic term of the equation of motion describing the fission process. In terms of the deformation parameter $`q`$ the mass is given, in the incompressible fluid approximation , as $$M=\pi \rho _0\underset{z_0}{\overset{z_0}{}}[\rho _s^2(z)A^2(z)+\frac{1}{2}B^2(z)]𝑑z,$$ $`(A7)`$ where $$A(z)=\frac{1}{\rho _s^2(z)}\frac{}{q}\underset{z}{\overset{z_0}{}}\rho _s^2(z^{})𝑑z^{},$$ $`(A8)`$ and $$B(z)=\frac{1}{2}\frac{\rho _s^2}{z}A+\frac{\rho _s^2}{q}$$ $`(A9)`$ and where $`\rho _0`$=0.17 fm<sup>-3</sup> is the matter density. The friction coefficient associated with the collective coordinate $`q`$ is calculated in the framework of the wall and window model . The wall contribution, which is the dominant part of the friction parameter $`\gamma `$ associated with the fission mode, is given by $$\gamma =\frac{\pi }{2}\rho _0\overline{v}\underset{z_0}{\overset{z_0}{}}\frac{\left(\frac{\rho _s^2}{q}\right)^2}{\sqrt{\rho _s^2(z)+\frac{1}{4}\left(\frac{\rho _s^2}{z}\right)^2}}𝑑z.$$ $`(A10)`$ where the average velocity $`\overline{v}`$ of the nucleons inside the nucleus is at zero temperature defined as $$\frac{\overline{v}}{c}=\frac{\overline{p}}{mc}=\frac{3}{4}\frac{\mathrm{}}{mc}(3\pi ^2\rho _0)^{1/3}$$ $`(A11)`$ with the Fermi momentum $`p_F=\mathrm{}k_F=\mathrm{}(3\pi ^2\rho _0)^{1/3}`$. Appendix B: Comparison between the microscopic semi-classical formulation of emission rates with the Weisskopf formula In Ref. , a formulation of emission rates has been given which is based on the picture of a Fermi gas of nucleons at temperature $`T`$ being the confined to a deformed rotating square well. In this appendix we shall show that in the special case of a spherical well, the model of Ref. becomes equivalent to Weisskopf’s emission rate formula . We consider the example of neutron emission. According to Ref. , the total number $`n`$ of neutrons emitted per time unit is given by $$n=\underset{\mathrm{\Sigma }}{}𝑑\sigma d^3pf_n(\stackrel{}{r}_0,\stackrel{}{p})v_{}(\stackrel{}{r}_0)w_0^{cl}(v_{}(\stackrel{}{r}_0)),$$ $`(B1)`$ where $`d\sigma `$ is the infinitesimal element of the surface $`\mathrm{\Sigma }`$ at the surface point $`\stackrel{}{r}_0`$. The momentum $`\stackrel{}{p}`$ and the velocity $`\stackrel{}{v}`$ of a neutron are related by the equation $$\stackrel{}{p}=m\stackrel{}{v}+m\stackrel{}{\omega }\times \stackrel{}{r},$$ $`(B2)`$ where $`\stackrel{}{r}`$ is the position vector, $`m`$ is the mass of the neutron, and $`\stackrel{}{\omega }`$ is the angular velocity of rotational motion of the nucleus. In the case of a spherical nucleus, a collective rotational motion is not possible. Consequently, we may put $`\stackrel{}{\omega }=0`$ without loss of generality. The Wigner function $`f_n(\stackrel{}{r}_0,\stackrel{}{p})`$ then describes a gas of fermions confined to a spherical well of depth $`V_0`$ at a temperature $`T`$. It is given by the expression $$f_n(\stackrel{}{r},\stackrel{}{p})=\frac{2}{h^3}\frac{\theta (\stackrel{}{r})}{1+\mathrm{exp}\left[\frac{1}{T}\left(\frac{p^2}{2m}V_0\mu _n\right)\right]},$$ $`(B3)`$ where $`\theta (\stackrel{}{r})`$ is a step function defined to be 1 for position vectors $`\stackrel{}{r}`$ of points inside the volume $`\mathrm{\Omega }`$ enclosed by the surface $`\mathrm{\Sigma }`$ $$\theta (\stackrel{}{r})=\{\begin{array}{cc}1\hfill & \mathrm{for}\stackrel{}{r}\mathrm{\Omega }\mathrm{including}\mathrm{\Sigma }\hfill \\ 0\hfill & \mathrm{otherwise}.\hfill \end{array}$$ $`(B4)`$ The parameter $`\mu _n`$ is the chemical potential of the neutrons. The last factor $`w_0^{cl}`$ in formula (B1) is the semi-classical transmission coefficient for a neutron hitting the surface $`\mathrm{\Sigma }`$ at the point $`\stackrel{}{r}_0`$ with a velocity $`v_{}(\stackrel{}{r}_0)`$ perpendicular to the surface. For a spherical surface $`\mathrm{\Sigma }`$ the unit vector normal to the surface at the surface point $`\stackrel{}{r}_0`$ is given by the radial unit vector $$\stackrel{}{n}_\mathrm{\Sigma }(\stackrel{}{r}_0)=\frac{\stackrel{}{r}_0}{R_0},$$ $`(B5)`$ where $`R_0:=|\stackrel{}{r}_0|`$ is the radius of the square well potential. In Weisskopf’s theory of evaporation, the total number of emitted particles of a given sort is represented as an integral over the energy $`\epsilon `$ of the emitted particles . Consequently, we would like to rewrite the formula (B1) as an integral over the variable $$\epsilon =\frac{p^2}{2m}V_0$$ $`(B6)`$ performing at the same time the remaining integrations in a closed form. For this purpose let us first introduce polar coordinates ($`p`$, $`\theta _p`$, $`\phi _p`$) in momentum space choosing the surface unit vector $`\stackrel{}{n}_\mathrm{\Sigma }(\stackrel{}{r}_0)`$ as the polar axis $`p_\xi =p\mathrm{sin}\theta _p\mathrm{cos}\phi _p,`$ $`p_\eta =p\mathrm{sin}\theta _p\mathrm{sin}\phi _p,`$ $`p_\zeta =p\mathrm{cos}\theta _p.`$ In these coordinates the perpendicular velocity $`v_{}(\stackrel{}{r}_0)`$ is given by $$v_{}(\stackrel{}{r}_0)=\frac{p\mathrm{cos}\theta _p}{m}.$$ $`(B7)`$ The transmission coefficient $`w_0^{cl}`$ can only be unequal to zero, if the velocity $`v_{}(\stackrel{}{r}_0)`$ is positive. Thus the polar angle $`\theta _p`$ can be restricted to the interval $`0\theta _p\pi /2`$. The total neutron yield per time unit is thus given by $$n=\frac{4\pi }{h^3}\underset{\mathrm{\Sigma }}{}𝑑\sigma \underset{0}{\overset{\mathrm{}}{}}𝑑p\underset{0}{\overset{\frac{\pi }{2}}{}}𝑑\theta _p\frac{p^3\mathrm{cos}\theta _p\mathrm{sin}\theta _p}{m}w_0^{cl}\left(\frac{p\mathrm{cos}\theta _p}{m}\right)\left[e^{\frac{1}{T}\left(\frac{p^2}{2m}V_0\mu _n\right)}+1\right]^1.$$ $`(B8)`$ The surface integral yields for factor $`4\pi R_0^2`$. Instead of integration variables $`p`$ and $`\theta _p`$, we introduce the energy $`\epsilon `$ and the orbital angular momentum $`l`$ of the neutron impinging on the surface $$\epsilon =\frac{p^2}{2m}V_0,$$ $`(B9)`$ $$l=R_0p\mathrm{sin}\theta _p.$$ $`(B10)`$ The Jacobian $`D`$ of the transformation $$D=\left|\begin{array}{cc}\frac{p}{\epsilon }\hfill & \frac{\theta _p}{\epsilon }\hfill \\ & \\ \frac{p}{l}\hfill & \frac{\theta _p}{l}\hfill \end{array}\right|$$ $`(B11)`$ is obtained from the inverse transformation $$\{\begin{array}{c}p=\sqrt{2m(\epsilon +V_0)}\hfill \\ \theta _p=\mathrm{arcsin}\left(\frac{l}{R_0p}\right)=\mathrm{arcsin}\left(\frac{l}{R_0\sqrt{2m(\epsilon +V_0)}}\right)\hfill \end{array}$$ $`(B12)`$ in the form $$D=\frac{\sqrt{2m/(\epsilon +V_0)}}{2\sqrt{2mR_0^2(\epsilon +V_0)l^2}}.$$ $`(B13)`$ In formulating the integration limits for the new variables $`\epsilon `$ and $`l`$ we must take into account the condition $$\frac{l^2}{2mR_0^2}\frac{p^2}{2m}=\epsilon +V_0,$$ $`(B14)`$ which implies that the total kinetic energy of the neutron can not be smaller than its rotational part. The factor $`p^3\mathrm{cos}\theta _p\mathrm{sin}\theta _p`$ can be written in terms of the new variables as follows $$p^3\mathrm{cos}\theta _p\mathrm{sin}\theta _p=2m(\epsilon +V_0)\sqrt{1\frac{l^2}{2mR_0^2(\epsilon +V_0)}}\frac{l}{R_0}.$$ $`(B15)`$ Making use the Eqs. (B13) to (B15) we obtain for the r.h.s. of Eq. (B8): $$\begin{array}{ccc}n& =& 2\pi \frac{2}{h^3}\mathrm{\hspace{0.17em}4}\pi R_0^2\underset{V_0}{\overset{\mathrm{}}{}}d\epsilon \underset{0}{\overset{\sqrt{2mR_0^2(\epsilon +V_0)}}{}}dl\frac{1}{2}\sqrt{\frac{2m}{(\epsilon +V_0)}}\\ & & \\ & & \frac{2(\epsilon +V_0)}{\sqrt{2mR_0^2(\epsilon +V_0)l^2}}\sqrt{1\frac{l^2}{2mR_0^2(\epsilon +V_0)}}\frac{l}{R_0}\frac{w_0^{cl}(\epsilon ,l)}{\left[e^{(\epsilon \mu _n)/T}+1\right]},\end{array}$$ $`(B16)`$ or $$n=\frac{(4\pi )^2}{h^3}\underset{V_0}{\overset{\mathrm{}}{}}𝑑\epsilon \underset{0}{\overset{\sqrt{2mR_0^2(\epsilon +V_0)}}{}}𝑑ll\frac{w_0^{cl}(\epsilon ,l)}{\left[e^{(\epsilon \mu _n)/T}+1\right]}.$$ $`(B17)`$ In Eq. (B17), the transmission coefficient $`w_0^{cl}(\epsilon ,l)`$ is obtained from $`w_0^{cl}(v_{})`$ by expressing the argument $`v_{}`$ by the variables $`\epsilon `$ and $`l`$. $$v_{}(\epsilon ,l)=\frac{\sqrt{2mR_0^2(\epsilon +V_0)l^2}}{mR_0}.$$ $`(B18)`$ Introducing the dimensionless angular momentum $`\lambda `$ instead of the variable $`l`$ $$l=\mathrm{}\lambda ,$$ $`(B19)`$ the formula (B17) takes the form $$n=\underset{V_0}{\overset{\mathrm{}}{}}𝑑\epsilon \underset{0}{\overset{\lambda _{\mathrm{max}}(\epsilon )}{}}𝑑\lambda \frac{^2n}{\epsilon \lambda },$$ $`(B20)`$ with $$\frac{^2n}{\epsilon \lambda }=\frac{4}{h}\frac{\lambda w^{cl}(\epsilon ,\lambda )}{\left[e^{(\epsilon \mu _n)/T}+1\right]}$$ $`(B21)`$ and $$\lambda _{\mathrm{max}}(\epsilon ):=\frac{1}{\mathrm{}}\sqrt{2mR_0^2(\epsilon +V_0)}.$$ $`(B22)`$ We wish to compare the result (B21) with the partial width $`\mathrm{\Gamma }_\nu ^{\alpha \beta }(E^{},\mathrm{\Lambda })`$ for the decay of a nucleus of excitation energy $`E^{}`$ and vanishing total angular momentum $`(\mathrm{\Lambda }=0)`$ by emission of a neutron of energy $`\epsilon _\alpha `$ and orbital angular momentum $`\lambda _\beta `$ as it is obtained from Weisskopf’s theory (Refs. , ) $$\mathrm{\Gamma }_n^{\alpha \beta }=\frac{2(2\lambda _\beta +1)}{h\rho (E^{},\mathrm{\Lambda }=0)}\rho _R(E_R^{},\mathrm{\Lambda }_R)w_n(\epsilon _\alpha ,\lambda _\beta ).$$ $`(B23)`$ In (B23), $`\rho (E^{},\mathrm{\Lambda })`$ and $`\rho _R(E_R^{},\mathrm{\Lambda }_R)`$ are the level densities for the mother and daughter nucleus, resp., which depend on the excitation energy $`E^{}(E_R^{})`$ and the total angular momentum $`\mathrm{\Lambda }(\mathrm{\Lambda }_R)`$ of the mother (daughter) nucleus. The quantity $`w_n(\epsilon _\alpha ,\lambda _\beta )`$ is the transmission factor for a neutron having the energy $`\epsilon _\alpha `$ and the orbital angular momentum $`\lambda _\beta `$. The factor 2 $`(2\lambda _\beta +1)`$, represents the product of the degeneracy factors of the emitted neutron and of the residual daughter nucleus. We note that, in the derivation of Eq. (B23) and also in our model, effects of the spin-orbit coupling on the emission probability are neglected. We now use the level densities obtained in the Fermi gas model: $$\rho _R(E_R^{},0)=\left(\frac{\mathrm{}^2}{2J}\right)^{3/2}\sqrt{a}\frac{\mathrm{exp}(2\sqrt{aE_R^{}})}{12E_R^2},$$ $`(B24)`$ $$\rho (E^{},0)=\left(\frac{\mathrm{}^2}{2J}\right)^{3/2}\sqrt{a}\frac{\mathrm{exp}(2\sqrt{aE^{}})}{12E^2}.$$ $`(B25)`$ Here $`J`$ is the rigid body moment of inertia of the nucleus which is assumed to be the same for the mother and daughter nucleus, and ”$`a`$” is the level density parameter. The excitation energies $`E^{}`$ and $`E_R^{}`$ are related by the energy conservation $$E^{}=E_R^{}+\epsilon _\alpha \mu _n.$$ $`(B26)`$ The ratio of the level densities is thus given by $$\begin{array}{c}\frac{\rho _R}{\rho }=\left(\mathrm{exp}\left[2\sqrt{aE_R^{}}2\sqrt{aE^{}}\right]\right)\left(\frac{E^{}}{E_R^{}}\right)^2\hfill \\ \\ \frac{\rho _R}{\rho }=\left(\mathrm{exp}\left[2\sqrt{aE^{}}\left(\sqrt{1\frac{\epsilon _\alpha \mu _n}{E^{}}}1\right)\right]\right)\left(1\frac{\epsilon _\alpha \mu _n}{E^{}}\right)^2.\hfill \end{array}$$ $`(B27)`$ Assuming that the ratio $`(\epsilon _\alpha \mu _n)/E^{}`$ is much smaller than 1 $$\frac{\epsilon _\alpha \mu _n}{E^{}}<<1$$ $`(B28)`$ and using the relation between the excitation energy $`E^{}`$ and the temperature $`T`$ as given by the Fermi gas model $$E^{}=aT^2$$ $`(B29)`$ we obtain for the ratio of the level densities $$\frac{\rho (E_R^{})}{\rho (E^{})}\mathrm{exp}\left(\frac{\epsilon _\alpha \mu _n}{T}\right).$$ $`(B30)`$ Using (B30) we obtain for the neutron width (B23) $$\mathrm{\Gamma }_n^{\alpha \beta }\frac{2(2\lambda _\beta +1)}{h}e^{\frac{\epsilon _\alpha \mu _n}{T}}w_n(\epsilon _\alpha ,\lambda _\beta ).$$ $`(B31)`$ This form of Weisskopf’s general formula agrees indeed with our result (B21) if the transmission factor $`w_n`$ is calculated classically ($`w_n=w_0^{cl}`$), if the orbital angular momentum $`\lambda _\beta `$ is large enough so that $$2\lambda _\beta +12\lambda _\beta $$ $`(B32)`$ and if $$\mathrm{exp}\left(\frac{\epsilon _\alpha \mu _n}{T}\right)>>1,$$ $`(B33)`$ so that the Fermi occupation factor in (B21) becomes equivalent to the Boltzmann factor $$\frac{1}{e^{\frac{\epsilon _\alpha \mu _n}{T}}+1}e^{\frac{\epsilon _\alpha \mu _n}{T}}.$$ $`(B34)`$ We note that the approximation (B34) is already used for obtaining the level density formula (B25) from the Fermi-gas model. We would like to comment that the well-known formula (B23) is already a slight extension of the original result of Weisskopf published in Ref. . There, the transmission factor $`w_n`$ is expressed in terms of the cross-section for the absorption of a neutron by the daughter nucleus using detailed balance. In this form, Weisskopf’s result is of very general validity. The difficulty with it is only that detailed balance relates the transmission factor $`w_n`$ to the cross-section for absorption of a neutron by a nucleus with excitation energy $`E_R^{}`$ and angular momentum $`\mathrm{\Lambda }_R`$. This absorption cross-section can, of course, not be measured. Therefore, one needs a model to calculate the transmission factor. The considerations in this appendix demonstrate at the same time that it is meaningful to replace the purely classical transmission factor $`w_0^{cl}`$ $$w_0^{cl}(\epsilon ,l)=\theta _0\left(\epsilon +V_0\frac{l^2}{2MR_0^2}\right)$$ $`(B35)`$ by a quantum-mechanical barrier penetration factor as it is done in all our numerical calculations. Table captions: 1. Multiplicities of the prefission particles emitted by: <sup>126</sup>Ba <sup>188</sup>Pt and <sup>266,272,278</sup>110 at different excitation energies. The theoretical estimates for the compound nuclei with Z=110 are done for two values of the friction force. In the first rows are the data evaluated with the friction reduced by 50% while those in the second row correspond to the standard wall-and-window friction. Figure captions: 1. The differential fusion (solid lines) and fission (bars) cross sections for the compound nucleus <sup>126</sup>Ba for different entrance channel reactions. 2. The deformation potential $`V_{\mathrm{fiss}}`$ for <sup>126</sup>Ba as a function of the ”fission variable” $`R_{12}/R_0`$ from the ground state deformation up to the scission point. The different curves correspond to the different values of the angular momentum $`L`$. The temperature of <sup>126</sup>Ba is fixed at $`T`$=1.6 MeV. 3. The same as in Fig. 2, but now the different curves correspond to the different temperatures $`T`$. All plots are done for $`L`$=30$`\mathrm{}`$. The different curves are shifted vertically in order to make the relative changes more visible. 4. The temperature of the compound nucleus <sup>126</sup>Ba at the different excitation energies as a function of angular momentum. 5. The multiplicities of prefission particles as a function of excitation energy of the compound nucleus <sup>126</sup>Ba obtained with the entrance channel <sup>28</sup>Si + <sup>98</sup>Mo. 6. The same as in Fig. 5 but for the reaction <sup>19</sup>F + <sup>107</sup>Ag. 7. The same as in Fig. 1 but for the compound nucleus <sup>188</sup>Pt. 8. The same as in Fig. 2 but for the compound nucleus <sup>188</sup>Pt. 9. The same as in Fig. 3 but for the compound nucleus <sup>188</sup>Pt. 10. The same as in Fig. 1 but for the compound nucleus <sup>272</sup>110. 11. The same as in Fig. 2 but for the compound nucleus <sup>266</sup>110 and temperature $`T`$=0. 12. The experimental (points with error bars) and theoretical multiplicities of prefission neutrons as a function of the excitation energy of the compound nuclei with $`Z=`$110 obtained in four different reactions. Table 1 | CN | Reaction | $`E_{lab}`$ | $`E^{}`$ | $`M_n^{exp}`$ | $`\delta M_n^{exp}`$ | $`M_n`$ | $`M_p`$ | $`M_\alpha `$ | | --- | --- | --- | --- | --- | --- | --- | --- | --- | | | | MeV | MeV | - | - | - | - | - | | | | 204.0 | 131.7 | 2.52 | 0.12 | 2.29 | 0.03 | 0.79 | | | <sup>28</sup>Si + <sup>98</sup>Mo | 187.2 | 118.5 | 2.01 | 0.13 | 1.71 | 0.00 | 0.09 | | <sup>126</sup>Ba | | 165.8 | 101.4 | 1.32 | 0.09 | 1.83 | 0.00 | 0.04 | | | | 142.8 | 84.1 | - | - | 0.27 | 0.04 | 0.88 | | | <sup>19</sup>F + <sup>107</sup>Ag | 147.8 | 118.5 | 1.85 | 0.11 | 1.99 | 0.00 | 0.16 | | | | 128.0 | 101.5 | 1.31 | 0.17 | 1.80 | 0.01 | 0.06 | | | <sup>34</sup>S + <sup>154</sup>Sm | 202.6 | 100.0 | 4.5 | 0.7 | 3.52 | 0.00 | 0.10 | | <sup>188</sup>Pt | | 159.8 | 66.5 | 2.5 | 0.7 | 1.10 | 0.00 | 0.00 | | | <sup>16</sup>O + <sup>172</sup>Yb | 137.6 | 99.7 | 5.4 | 0.6 | 3.79 | 0.00 | 0.06 | | | | 513.9 | 185.9 | 7.83 | 0.46 | 8.66 | 0.26 | 1.04 | | | | | | | | 9.72 | 0.23 | 0.95 | | | | 486.6 | 164.6 | 7.35 | 0.50 | 7.65 | 0.20 | 0.93 | | | | | | | | 8.56 | 0.21 | 0.79 | | <sup>266</sup>110 | <sup>58</sup>Ni + <sup>208</sup>Pb | 461.1 | 144.6 | 6.17 | 0.48 | 6.50 | 0.16 | 0.81 | | | | | | | | 7.30 | 0.17 | 0.73 | | | | 436.2 | 125.1 | 4.74 | 0.49 | 5.45 | 0.10 | 0.63 | | | | | | | | 6.21 | 0.11 | 0.58 | | | | 410.1 | 104.7 | 4.41 | 0.41 | 4.35 | 0.07 | 0.48 | | | | | | | | 4.94 | 0.07 | 0.48 | | | | 377.0 | 78.8 | 2.94 | 0.36 | 2.77 | 0.03 | 0.30 | | | | | | | | 3.23 | 0.03 | 0.33 | | | | 472.3 | 138.3 | 5.98 | 0.43 | 7.49 | 0.08 | 0.42 | | | | | | | | 8.25 | 0.08 | 0.34 | | | | 449.9 | 121.1 | 5.52 | 0.38 | 6.40 | 0.05 | 0.30 | | | | | | | | 7.12 | 0.06 | 0.28 | | | <sup>64</sup>Ni + <sup>208</sup>Pb | 423.0 | 100.6 | 5.13 | 0.33 | 5.09 | 0.02 | 0.20 | | | | | | | | 5.58 | 0.04 | 0.20 | | | | 403.9 | 85.9 | 3.54 | 0.31 | 3.94 | 0.01 | 0.15 | | | | | | | | 4.49 | 0.01 | 0.12 | | <sup>272</sup>110 | | 377.0 | 65.3 | 3.03 | 0.32 | 2.27 | 0.00 | 0.10 | | | | | | | | 2.83 | 0.00 | 0.08 | | | | 351.2 | 166.3 | 8.40 | 0.53 | 9.23 | 0.13 | 0.53 | | | | | | | | 9.95 | 0.15 | 0.48 | | | <sup>40</sup>Ca + <sup>232</sup>Th | 298.4 | 121.3 | 5.81 | 0.50 | 6.40 | 0.05 | 0.30 | | | | | | | | 7.12 | 0.06 | 0.28 | | | | 250.4 | 80.3 | 3.35 | 0.34 | 3.56 | 0.01 | 0.14 | | | | | | | | 4.04 | 0.01 | 0.12 | | | | 299.6 | 127.2 | 5.78 | 0.51 | 7.89 | 0.03 | 0.14 | | | | | | | | 8.48 | 0.04 | 0.15 | | <sup>278</sup>110 | <sup>40</sup>Ar + <sup>238</sup>U | 280.4 | 110.7 | 4.96 | 0.55 | 6.69 | 0.00 | 0.12 | | | | | | | | 7.23 | 0.03 | 0.13 | | | | 258.0 | 91.5 | 4.22 | 0.44 | 5.08 | 0.00 | 0.09 | | | | | | | | 5.59 | 0.00 | 0.09 |
warning/0001/hep-ex0001038.html
ar5iv
text
# A partial wave analysis of the 𝜋⁰⁢𝜋⁰ system produced in 𝜋⁻⁢𝑝 charge exchange collisions ## I Introduction This paper reports on a high-statistics partial wave analysis (PWA) of the $`\pi ^0\pi ^0`$ system produced in the charge exchange reaction: $`\pi ^{}p\pi ^0\pi ^0n`$ at an incident momentum of $`18.3GeV/c`$ using data taken by experiment E852 at Brookhaven National Lab (BNL). The PWA was performed over the $`m_{\pi ^0\pi ^0}`$ mass range from near threshold ($`2m_{\pi ^0}`$) to as high as $`2.2GeV/c^2`$ in $`0.04GeV/c^2`$ mass bins and in various bins in momentum-transfer-squared $`t=\left|p_\pi p_{\pi \pi }\right|^2=\left|p_np_p\right|^2`$. The mass and $`\left|t\right|`$ dependence of $`\pi \pi `$ production in $`\pi ^{}`$\- induced reactions with one pion exchange (OPE) provides information on the process $`\pi \pi \pi \pi `$, involving the scattering of the lightest hadrons . The extraction of $`\pi \pi \pi \pi `$ amplitudes is, however, complicated by the presence of production mechanisms other than OPE . The $`|t|`$ and $`m_{\pi \pi }`$-dependence of the partial wave amplitudes and their relative phases, the focus of this paper, provide information on these mechanisms and the necessary input for future $`\pi \pi `$ scattering studies. The study of the $`\pi \pi `$ system also bears on current issues in the spectroscopy of conventional $`q\overline{q}`$ mesons and non-$`q\overline{q}`$ mesons such as glueballs or mesonic molecules. In particular, the isoscalar scalar and tensor sectors have more states than can be accommodated within the conventional $`q\overline{q}`$ model. A recent review of light meson spectroscopy includes a summary of the current experimental situation in these sectors. Non-$`q\overline{q}`$ candidates include the poorly understood $`f_0(980)`$ and the glueball candidates $`f_0(1500)`$ and $`f_J(1710)`$, all of which couple to the $`\pi \pi `$ system . Information about the masses,widths, and decay modes of these states, along with knowledge of their production mechanisms, as revealed by their $`|t|`$ dependences, will help in unraveling their substructure . A complete understanding of these states requires corresponding information from $`\eta \eta `$ and $`K\overline{K}`$ final states as well. This paper presents information which may be used in such a program. The $`J^{PC}`$ of the $`\pi ^0\pi ^0`$ system must have J even with both P and C positive. The isospin must also be even $`(\text{I = 0}`$ or $`\text{I = 2})`$ for $`\pi ^0\pi ^0`$. The $`\pi ^0\pi ^0`$ system is thus particularly attractive for investigation of scalar and tensor states as the PWA is simplified without the presence of odd angular momenta. The $`\pi ^{}p\pi ^0\pi ^0n`$ reaction has been studied in experiments with incident $`\pi ^{}`$ momenta of $`9GeV/c`$ , $`25GeV/c`$ , $`38GeV/c`$ and $`100GeV/c`$ . The combined information from these experiments can be used to provide information on how cross sections of produced states and relative ratios of partial waves depend on center-of-mass energy. This paper is organized as follows: The experimental overview is presented in Section 2. Event reconstruction and data selection are described in Section 3, where the general features of the distributions in $`m_{\pi ^0\pi ^0}`$ and $`|t|`$ are also discussed. The details of the PWA formalism and results are given in Section 4. In Section 5 Regge-models are fitted to the results from Section 4. The conclusions are summarized in Section 6. ## II Experimental Overview The E852 apparatus was built around and included the Multi-Particle Spectrometer (MPS) at BNL. The data used for the analysis reported in this paper were collected in 1994 and 1995 using a beam of negatively charged particles of momentum $`18.3GeV/c`$. A 30-cm liquid hydrogen target was surrounded by a cylindrical drift chamber and an array of thallium-doped CsI crystals arranged in a barrel, all located inside the MPS dipole magnet. Drift chambers were used to track charged particles downstream of the target. Two proportional wire chambers (PWC’s), downstream of the target, were used in requiring specific charged particle multiplicities in the event trigger. A 3000-element lead glass detector (LGD) measured the energies and positions of photons in the forward direction. The dimensions of the LGD matched the downstream aperture of the MPS magnet. Photons missing the LGD were detected by the CsI array or by a lead/scintillator sandwich array (DEA) arranged in a picture frame downstream of the target with an aperture to allow for the passage of charged particles. The first level trigger required that the unscattered or elastically scattered beam not enter an arrangement of two small beam-veto scintillation counters located in front of the LGD. The next level of trigger required that there be no signal in the DEA and no charged particles recorded in the cylindrical drift chamber surrounding the target or in the PWC’s (an all-neutral trigger). In the 1994 run, all layers of the cylindrical drift chamber were used in the trigger requirement, whereas in the 1995 run, only the outer layer was used. A common off-line analysis criterion required no hits in the cylindrical drift chamber. The final trigger requirement was a minimum deposition of electromagnetic energy in the LGD. The LGD is central to this analysis and it is described in detail in reference . The LGD was initially calibrated by moving each module into a monoenergetic electron beam. Further calibration was performed by adjusting the calibration constant for each module until the width of the $`\pi ^0`$ and $`\eta `$ peaks in the $`\gamma \gamma `$ effective mass distribution was minimized. The calibration constants were also used for a trigger processor which did a digital calculation of energy deposited in the LGD and the effective mass of photons striking the LGD. A laser-based monitoring system allowed for tracking the gains of individual modules. Studies were made of various algorithms for finding cluster of energies deposited by photons including issues of photon-to-photon separation and position finding resolution. These are also described in reference . ## III Event Reconstruction and Data Selection The combined data sets taken in 1994 and 1995 contain approximately 70 million all-neutral triggered events. Of these events, approximately 13 million were found to have four photons in the LGD. The plot of di-photon effective masses for all possible pairings of photons is shown in the scatterplot of figure 1(a) and the projection is shown in figure 1(b). Events consistent with the production of two $`\pi ^0`$’s dominate the scatterplot. The $`\pi ^0`$ mass resolution is 17 $`MeV/c^2`$. The sample of 847,460 $`\pi ^{}p\pi ^0\pi ^0n`$ events was selected from the 13 million four photon events by imposing various analysis criteria. It was required that no charged particles were registered in the MPS drift chambers or the cylindrical drift chamber surrounding the liquid hydrogen target. Any event with a photon within $`8cm`$ of the center of the beam hole or the outer edge of the LGD was removed. The $`\chi ^2`$ returned from kinematic fitting to the $`\pi p\pi ^0\pi ^0n`$ reaction hypothesis was required to be less than 9.8 (95% C.L. for a three-constraint fit). A further demand was that none of the other final state hypotheses considered ($`\eta \pi ^0n`$, $`\eta \eta n`$) had a better $`\chi ^2`$. The final criterion was that the CsI detector registered less than 20 MeV, a cut which eliminated events with low-energy $`\pi ^0`$’s. The $`\pi ^0\pi ^0`$ mass resolution improves from 24 $`MeV/c^2`$ to 16 $`MeV/c^2`$ at the mass of the $`K_s^0`$ after kinematic fitting. Background studies were also carried out. By selecting events in a given four photon effective mass region and fitting the associated scatter plot of di-photon effective mass pairings (similar to figure 1), the background of non-$`\pi ^0\pi ^0`$events under the signal was found to be very small. Typical signal to noise ratios determined by these studies are in the range of 50:1. Monte Carlo studies indicate that combinatoric background from mis-pairing the reconstructed photons is a few percent below $`m_{\pi \pi }0.5GeV/c^2`$ and non-existent at higher masses. These studies are described in more detail in reference . The distribution in missing-mass-squared, recoiling against the four photons, for events with a successful kinematic fit to the reaction $`\pi ^{}p\pi ^0\pi ^0n`$ is shown in figure 2. The missing-mass-squared is determined from photon position and energy information before kinematic fitting and the distribution peaks near the square of the neutron mass. The distribution in $`\pi ^0\pi ^0`$ effective mass is shown in figure 3. The spectrum is dominated by the $`f_2(1270)`$ resonance and a broad enhancement at low $`\pi ^0\pi ^0`$ mass (from threshold to about $`1.0GeV/c^2`$). There is also a small $`K_S^0\pi ^0\pi ^0`$ signal present, despite the requirement that the deposited energy in the CsI detector not exceed $`20MeV`$. This CsI energy cut reduces a substantial fraction of $`K_S^0`$ events but other reactions producing $`K_S^0`$ can avoid deposition of energy in the CsI detector. By correlating the observed yields of $`K_S^0`$ and $`f_2(1270)`$ mesons, for samples with and without the CsI detector energy cuts, with cross sections for $`f_2(1270)`$ production and associated $`K_S^0`$ production ($`\pi ^{}pK_S^0\mathrm{\Lambda }(\mathrm{\Sigma }^0)`$) measured in other experiments, we estimate an overall CsI detector inefficiency of $`5\%`$. These studies also indicate that the background level of non-neutron events under the $`f_2(1270)`$ is approximately $`1\%`$. Another feature of the spectrum is the dip at $`1.0GeV/c^2`$, which will be seen to be due to the interference of a narrow resonance, the $`f_0(980)`$, with a broad $`\pi ^0\pi ^0`$ enhancement. The distribution in $`|t|`$, shown in figure 3, is not characterized by a single exponential, suggesting more than one production mechanism. The curve is a fit of this distribution to a sum of two exponentials: $`dN/dt=ae^{b|t|}+ce^{d|t|}`$ where $`b=15.5\left(GeV/c\right)^2`$ and $`d=3.7\left(GeV/c\right)^2`$. Based on this structure, we initially examine the $`\pi ^0\pi ^0`$ effective mass spectra in four bins in $`|t|`$ as shown in figure 4. The $`t`$-dependence of the $`S`$, $`D_0`$, and $`D_+`$ partial waves is later investigated in a set of partial wave fits more finely binned in $`|t|`$. An inspection of figure 4 reveals striking differences in the $`\pi ^0\pi ^0`$ mass spectra associated with the four bins in $`|t|`$. For example, the low-mass structure which dominates in figure 4a is much less prominent in figure 4d. The dip associated with the $`f_0(980)`$ resonance in figure 4a becomes a bump in figure 4d. These and other features are explored in more detail below in the discussion of the PWA results. ## IV Partial Wave Analysis Partial wave analysis is used to extract production amplitudes (partial waves) from the observed decay angular distributions of the di-pion system. A process such as $`\pi ^{}p\pi ^0\pi ^0n`$, dominated by $`t`$-channel meson exchange, is simplest to analyze in the Gottfried-Jackson reference frame. The Gottfried-Jackson frame is defined as a right-handed coordinate system in the center of mass of the produced di-pion system with the $`zaxis`$ defined by the beam particle momentum and the $`yaxis`$ perpendicular to the plane defined by the beam and recoil neutron momenta. The decay angles ($`\theta ,\varphi `$) are determined for one of the produced $`\pi ^0`$ momenta. At fixed beam momentum, an event is fully specified by $`(m_{\pi \pi },t,\theta ,\varphi )`$. The data are binned in $`m_{\pi \pi }`$ and $`t`$ and the production amplitudes, and their relative phases, are extracted from the accumulated angular distributions using an extended maximum likelihood fit to the distributions in ($`\theta `$,$`\varphi `$) . The naming convention for the partial waves is summarized in Table I. The explicit form of the angular distribution $`I(\theta ,\varphi )`$ fitted to the data in a given mass and momentum transfer range in this analysis is given by $`I(\theta ,\varphi )=`$ $`\left|S+\sqrt{5}D_0P_2^0(\mathrm{cos}\theta )\sqrt{{\displaystyle \frac{5}{3}}}D_{}P_2^1(\mathrm{cos}\theta )\mathrm{cos}\varphi +\sqrt{9}G_0P_4^0(\mathrm{cos}\theta )\right|^2`$ (1) $`+\left|\sqrt{{\displaystyle \frac{5}{3}}}D_+P_2^1(\mathrm{cos}\theta )\mathrm{sin}\varphi \right|^2`$ (2) where $`P_l^m(\theta )`$ are the associated Legendre polynomials . As summarized in Table I, the $`D_+`$ wave is produced by the exchange of a particle with natural parity ($`P=(1)^J`$). For production of a $`\pi \pi `$ system, the dominant natural parity exchange particle is the $`a_2`$ . The $`S`$, $`D_0`$, $`D_{}`$ and $`G_0`$ waves are produced by the exchange of a particle with unnatural parity ($`P=(1)^{J+1}`$). Again, for $`\pi \pi `$ production, the dominant unnatural parity exchange particles are the $`\pi `$ and the $`a_1`$ . ### A Ambiguities There are multiple discrete sets of partial wave amplitudes which can give rise to exactly the same angular distribution . It can be shown that in a partial wave fit with only $`S,D_0,D_,`$ and $`D_+`$ partial waves there are four sets of ambiguous partial wave amplitudes. The four sets can be divided into two groups with different partial wave intensities. Additionally, within each group, there is a sign ambiguity in the phases between the amplitudes. Normally there are two ambiguities; if one wave in each naturality is fixed *a priori*, e.g. set real or to some complex value for dynamical reasons, there is still an overall sign ambiguity. However, even this sign ambiguity could be fixed by the requirement, for example, of a resonant behavior in one of the waves. In general, there is an eightfold ambiguity for a $`\pi ^0\pi ^0`$ system containing L=0, 2 and 4. However, these ambiguities necessarily entail nonzero $`G_{}`$ and $`G_+`$ waves. In this paper we have assumed that these are negligibly small and searched for ambiguities with nonzero $`G_0`$ wave. We find no such ambiguities in our data. In the analysis of the $`\pi ^0\pi ^0`$ system, the physical solution can be selected by a combination of physical arguments (which will be given below) and the requirement that solutions be smoothly connected as a function of mass. This selection of the physical solution applies simultaneously to all intensities and phases. In what follows, the physical solution is plotted with solid symbols. The other solutions are plotted with open symbols and are presented for completeness. ### B Partial Wave Fits #### 1 Results for $`0.01<t<0.10GeV^2/c^2`$ The results of the partial wave decomposition are shown in figures 5 and 6. The partial wave intensities are shown in figure 5 and the phase differences in figure 6. The phase difference plots are shown above $`\pi ^0\pi ^0`$ masses of 0.8 $`GeV/c^2`$. Below that value, where one of the waves is very small, phase difference information is unreliable. As discussed in Section IV A, there is a two-fold ambiguity in the intensities. The threshold behavior ($`S`$-wave dominance) and the resonant behavior of the $`f_2(1270)`$ are used to select the physical solution. Furthermore, since the resonant structures of both the $`D_0`$ and $`D_{}`$ partial waves are due to the $`f_2(1270)`$, the relative phase between the $`D_0`$ and $`D_{}`$ partial waves should be constant and near $`\pm \pi `$ radians, according to the phase convention of . These assumptions allow the physical solution at low mass to be connected with the solutions at higher mass. Above approximately $`1.5GeV/c^2`$, the solutions become degenerate. The spin-4 $`G_0`$ partial wave is not included in the fit below $`1.4GeV/c^2`$. There are a number of key features observed in the physical solution. There is at least one broad enhancement in the $`S`$-wave intensity and a sharp dip in the $`S`$-wave intensity near $`1.0GeV/c^2`$ accompanied by rapid phase variation in the $`S`$ $`D_0`$ relative phase. There also exists a dip in the $`S`$-wave intensity near $`1.5GeV/c^2`$ accompanied by rapid phase variation in the $`S`$ $`D_0`$ relative phase. The $`f_2(1270)`$ is observed in the $`D_0,D_{}`$, and $`D_+`$ partial wave intensities, and the bump observed in the $`G_0`$ partial wave near 2.0 $`GeV/c^2`$ is consistent with the $`f_4(2040)`$. Finally, the $`D_0`$-wave intensity is larger than the $`D_{}`$-wave intensity or the $`D_+`$-wave intensity, consistent with the expectation that OPE should favor production of an m=0 wave for this low-$`|t|`$ region. A background term was not included in the PWA fits presented in this paper. A background term was included in some earlier fits where it was found that below about 1.0 $`GeV/c^2`$ it cannot be distinguished from the dominant $`S_0`$ wave and above 1.0 $`GeV/c^2`$, the fit forces the background term to zero. #### 2 Results for $`0.10<t<0.20GeV^2/c^2`$ The results of the partial wave analysis (Figures 7 and 8) in this region are qualitatively similar to the results in the $`0.01<t<0.10GeV^2/c^2`$ region. The same techniques are used to select the physical solution as in the previous region in $`|t|`$. The $`S`$-wave intensity contains at least one broad object and two dips. The $`f_2(1270)`$ is observed in all $`D`$-waves. An enhancement near 2.0 $`GeV/c^2`$ is again observed in the $`G_0`$ partial wave. More detailed comparisons with the results from the $`0.01<t<0.10GeV/c^2`$ region reveal the following differences: The ratio of the $`S`$-wave intensity to the $`D_0`$-wave intensity is smaller at larger $`|t|`$ and the ratio of the $`D_0`$-wave intensity to both the $`D_{}`$-wave and $`D_+`$-wave intensities is smaller at larger $`|t|`$. The ratio of the $`D_0`$-wave intensity to $`G_0`$-wave intensity does not change suggesting that the $`f_2(1270)`$ and the $`f_4(2040)`$ are produced by the same mechanism. #### 3 Results for $`0.20<t<0.40GeV^2/c^2`$ The change in slope for the $`|t|`$-distribution as seen in figure 3, indicates a change in production mechanism. This is reflected in the partial wave analysis as well (Figures 9 and 10). For this $`|t|`$ region, the $`G_0`$ partial wave is not required for an adequate description of the observed angular distributions and is therefore not included here or in the next higher $`|t|`$ region. The $`S`$-wave intensity has a different shape compared to that at smaller values of $`|t|`$. The $`D_+`$-wave intensity is approximately one-third as large as the $`D_0`$-wave intensity whereas at smaller momentum transfer it was approximately one-tenth as large. #### 4 Results for $`0.40<t<1.50GeV^2/c^2`$ The partial wave analysis results in the region $`0.40<t<1.50GeV^2/c^2`$ (Figures 11 and 12) are significantly different from results at smaller $`|t|`$. The bump observed in the mass plot (figure 4d) near $`1.0GeV/c^2`$ is found in the $`S`$-wave intensity. The $`D_+`$ partial wave is dominant (as opposed to the $`D_0`$ partial wave), indicating a shift from unnatural parity exchange processes at small $`|t|`$ to production via natural parity exchange at large $`|t|`$. #### 5 Fine $`|t|`$ Bin Fits The statistics of this experiment are sufficient to allow the region $`0.00<t<0.40GeV^2/c^2`$ to be analyzed in finer $`|t|`$ bins, nine in all, for masses up to approximately $`1.8GeV/c^2`$. The $`|t|`$ -dependence of the $`S`$-wave intensity may be summarized by noting that the ratio of the maxima in the intensities at approximately $`0.8GeV/c^2`$ and $`1.3GeV/c^2`$ decreases with increasing $`|t|`$, and the ratio of height of maximum intensity at approximately $`0.8GeV/c^2`$ to value of the intensity measured at $`0.98GeV/c^2`$ decreases. The lineshape of the $`f_2(1270)`$ in the $`D_0`$-wave intensity is largely independent of $`|t|`$. The $`SD_0`$ relative phase is $`|t|`$ dependent. The lineshape of the $`D_+`$-wave is also independent of $`|t|`$. More details of the $`|t|`$ dependence of the partial waves follow. The intensities of the individual partial-waves and phase differences as a function of mass for the nine bins in $`|t|`$ for $`0.00<t<0.40GeV^2/c^2`$ as well as for the $`|t|`$-bins presented in this paper are available on the World Wide Web . ### C Model Dependent Fits of the $`|t|`$\- Distributions The integrals of fitted relativistic Breit-Wigner functions over the peak regions of the $`D_0`$ and $`D_+`$-waves as a function of $`|t|`$ are shown in Figure 13. The dependences of these intensities on $`|t|`$ are fitted to functions given by Regge-exchange models. At low-$`|t|`$, the unnatural parity exchange $`D_0`$ partial wave is expected to be dominated by OPE. The Reggeized form for this contribution is given by $$\frac{d|D_0|}{d|t|}=N_{D_0}|\sqrt{t}e^{b_\pi t}(tm_{f_2}^2)^2\left(1+e^{i\pi \alpha (t)}\right)\mathrm{\Gamma }(\alpha _\pi (t))|^2,\alpha _\pi (t)=0.9(tm_\pi ^2)$$ (3) In this expression, the $`\sqrt{t}`$ factor is due to helicity-flip in the pion-nucleon coupling, and the polynomial dependence on $`t`$ arises from the $`f_2`$ coupling to $`\pi \pi `$ at the production vertex. The particular form of this dependence is due to the angular momentum barrier factor proportional to $`k^L`$ with $`L=2`$ and $`k`$ being the magnitude of the 3-momentum of the exchanged particle in the $`f_2`$ rest frame (Gottfried-Jackson frame), given by $`k^2=((m_{f_2}m_\pi )^2t)((m_{f_2}+m_\pi )^2t)/4m_{f_2}^2(m_{f_2}^2t)^2/4m_{f_2}^2`$. The slope, $`b_\pi `$, in the OPE form is $`4.08\pm 0.02/(GeV^2/c^2)`$. The systematic uncertainty in the slope of the $`\alpha _\pi (t)`$ Regge trajectory is $`\pm 0.1/(GeV^2/c^2)`$. As shown by Irving and Michael the natural parity exchange $`D_+`$ -wave is dominated by absorption of the pion exchange and may be parameterized in terms of a Regge cut in the nucleon helicity-flip amplitude $$C=g_ce^{b_ct}e^{\frac{1}{2}i\pi \alpha _C(t)}\left(\frac{p_L}{p_0}\right)^{\alpha _C(t)1},\alpha _C(t)=0.41t,g_c=\mathrm{0\; .84},b_c=3.89$$ (4) The nucleon-flip and non-flip $`a_2`$ exchange is then given by, $$A_f=g_a(t)e^{b_at}e^{\frac{1}{2}i\pi \alpha _{A_2}(t)}\left(\frac{p_L}{p_0}\right)^{\alpha _{A_2}(t)1},A_n=A_f\frac{r}{\sqrt{t}}$$ (5) respectively, with the parameters $`\alpha _{A_2}(t)=0.5+0.82t`$ and $`g_a=1.35`$, $`b_a=3.24`$, $`p_0=17.2GeV/c`$, and $`r=0.5`$ from and $`p_L=18.3GeV/c`$, the beam momentum for these data. The $`D_+`$-wave intensity is then fitted to $$\frac{d|D_+|^2}{d|t|}=N_{D_+}(|A_n|^2+|A_f+C|^2)$$ (6) For both forms the fitted functions are averaged over the $`|t|`$ bins shown in the plots. The plotted curves are calculated from the models without averaging. In figure 14 the peak value of the $`S`$-wave intensity near $`0.80GeV/c^2`$, the value of the $`S`$-wave intensity at $`0.98GeV/c^2`$ and the peak value of the $`S`$-wave intensity at approximately $`1.3GeV/c^2`$ as a function of $`|t|`$ are shown. A one-pion-exchange form similar to Equation 3, but with the $`tm_{f_2}^2`$ factor removed, is overlayed on these distributions. The Regge trajectory slope and exponential slope are fixed to the values found for the $`D_0`$-wave fit, and a one parameter fit is used to set the normalization. At small values of $`|t|`$ the OPE form qualitatively agrees with the data. The excess of events at higher $`|t|`$ in (b) and (c) is consistent with the existence of additional production mechanisms that are less strongly biased toward small momentum-transfer-squared production than is OPE. ## V Conclusions A partial wave analysis was carried out on a sample of 847,460 events of the reaction $`\pi ^{}p\pi ^0\pi ^0n`$ collected by experiment E852. The PWA was performed in $`0.04GeV/c^2`$-wide bins in di-pion mass ($`m_{\pi ^0\pi ^0}`$) and momentum-transfer-squared ($`|t|`$) from the incident $`\pi ^{}`$ to the outgoing $`\pi ^0\pi ^0`$ system. Coarse and fine binning in $`|t|`$ were used. Numerical values for the partial wave intensities and phases as a function of di-pion mass, for coarse and fine bins in $`|t|`$, are available on the World Wide Web . The $`f_2(1270)`$ meson is found to be produced by unnatural parity exchange at small values of $`|t|`$ and natural parity exchange at large values of $`|t|`$. The $`|t|`$ dependences of $`D_0`$-wave and $`D_+`$-wave intensities are consistent with Regge-exchange models. An enhancement in the $`G_0`$ wave consistent with the $`f_4(2050)`$ meson is observed in unnatural parity exchange at small momentum transfer. The shape of the $`S`$-wave intensity has a strong momentum transfer dependence. The presence of dips in the $`S`$-wave intensity near $`0.98`$ and $`1.5GeV/c^2`$, accompanied by rapid phase variations relative to the $`D_0`$-wave is consistent with similar observations reported in reference and in centrally produced $`\pi ^0\pi ^0`$ systems in 450 $`GeV/c`$ $`pp`$ collisions . The latter claims evidence for the $`f_0(980)`$ and $`f_0(1500)`$. At large momentum transfer, the $`f_0(980)`$ meson is observed as a bump in the $`S`$-wave intensity. The $`S`$-wave intensity in the peak near $`0.80GeV/c^2`$ is well-described by OPE. It should also be noted that the model of Anisovich et al predicts the presence of a dip in the $`|t|`$ distribution for this mass region near $`|t|0.07GeV^2/c^2`$ . In direct contradiction, no such dip is observed in this analysis. At higher masses the $`S`$-wave is adequately described by OPE only at small values of $`|t|`$. ## VI Acknowledgements We wish to thank the members of the MPS group at BNL as well as the staffs of the AGS, of BNL and of the various colaborating institutions. This work was supported in part by the U. S. Department of Energy, the National Science Foundation and the Russian State Committee for Science and Technology.
warning/0001/cond-mat0001297.html
ar5iv
text
# One dimensional model for doubly degenerate electrons ## I Introduction There has been much interests in the studies on correlated electrons in the presence of orbital degree of freedom. The orbital degree of freedom is relevant to many transitional metal oxides . It may be also relevant to some $`C_{60}`$ materials and samples of artificial quantum dot arrays . For a theoretical understanding of the observed unusual properties, a SU(4) theory describing spin systems with orbital degeneracy was proposed . There were also numerical and perturbative studies of 1-dimensional models for these systems. The ground-state phase diagrams for the system with a symmetry breaking of $`SU(4)SU(2)\times SU(2)`$ were discussed . The phase separation was recently observed in experiment. Along with the rapid developments in experiments where the metal ions has orbital degeneracy in addition to spin degeneracy, a theoretical study of such system by taking account of the kinetic terms caused by nearest neighbor hopping becomes important. In a recent letter the Hubbard model for electrons with orbital degeneracy was studied. It was shown that the model not only has an underling SU(4) symmetry of spin-orbital double but also has a hidden charge SU(4) symmetry. An extended Lieb-Mattis transformation which maps those two SU(4) generators into each other is given. On the basis of elementary degenerate perturbative theory, it was shown that the effective Hamiltonian with strong repulsive coupling at half-filling is equivalent to the Hamiltonian of the $`SO(6)`$ Heisenberg model, and that at quarter-filling is equivalent to the one of SU(4) Heisenberg model. Some features of the model in one dimension was also briefly described. In present paper, we study the one-dimensional Hubbard-like model with SU(4) symmetry for electrons with two-fold orbital degeneracy extensively. Its exact solution is formulated by means of Bethe ansatz if sites are assumed to be occupied by at most two electrons. The features of the ground state and excited states for repulsive coupling are shown. For finite $`N`$ number of electrons, the configurations of quantum numbers are given explicitly and the spectra of excitations are obtained by solving the Bethe-ansatz equation numerically. For infinite $`N`$, the ground state and various kind of low-lying excitations are obtained on the basis of thermodynamics limit. The paper is organized as follows. In next section we propose the model Hamiltonian with some interpretative remarks. Employing standard method we carefully formulated the first quantized version of the Hamiltonian. We make an allowed modification so that the Bethe-Yang ansatz is applicable to this model. A detailed formulation from the Bethe-ansatz wave function to the Bethe-ansatz equation is given. In Sec. III, we explicitly show how the quantum numbers in the Bethe-ansatz equation should be taken for the non-degenerated ground state. We calculate the ground-state energy and Fermi momentum numerically for different numbers of electrons. We also compare them for different coupling constants. In Sec. IV, we study the excited states extensively by analyzing the possible variations in the sequence of quantum numbers. We indicate in each case how the quantum numbers change from integers to half-integers (or vise versa) with respect to that of the ground state. Numerical results of energy-momentum spectra for each excitation are given there. In Sec. V, two special cases, weak and strong coupling are discussed. We are able to deduce several interesting properties from the Bethe-ansatz equation without solving it directly. In Sec. VI, we consider the thermodynamics limit. After giving some general formulae and expressions, we study the ground state and calculate the ground-state energy explicitly for strong coupling. In Sec. VII, we discuss low-lying excitations in the spin-orbital sector on the basis of thermodynamics limit. Both contributions of holes and 2-strings are taken into account. The singlet excitation and several multiplet excitations are obtained. In Sec. VIII, we discuss low-lying excitations in charge sector by thermodynamics limit. The holon-antiholon and holon-holon excitations are obtained. Sec. IX is a summary of the main results of the paper and some conclusive discussions. ## II The model and its exact solution We consider a Hubbard-like model for electrons with two-fold orbital degeneracy. The spin components are denoted by up $`()`$ and down $`()`$, the orbital components by top and bottom. The four possible states of electrons are $`|1>=|\begin{array}{c}\\ \end{array}>,`$ $`|2>=|\begin{array}{c}\\ \end{array}>,`$ (5) $`|3>=|\begin{array}{c}\\ \end{array}>,`$ $`|4>=|\begin{array}{c}\\ \end{array}>.`$ (10) We use $`1,2,3,`$ and $`4`$ to label these states from now on. The model Hamiltonian then reads $$H=t\underset{i,a}{}(C_{i,a}^+C_{i+1,a}+C_{i+1,a}^+C_{i,a})+U\underset{i,a<a^{}}{}n_{i,a}n_{i,a^{}}$$ (11) where $`i=1,2,\mathrm{},L`$ identify the lattice site, $`a,a^{}=1,2,3,4`$ specify the spin and orbital as defined in the above. The $`C_{ia}^+`$ creates an electron with spin-orbital component $`a`$ on site $`i`$, and $`n_{ia}:=C_{ia}^+C_{ia}`$ is the corresponding number operator at site $`i`$. Eq.(11) is the Hamiltonian for four-component systems, and there were various discussions on multi-component Hubbard model in one dimension . Whereas the physics that eq.(11) describes will be precise only when the representation space for the internal degree of freedom is specified . It is specified to the spin and orbital, and the site is assumed to be occupied by at most two electrons in our present model. It is convenient to consider the states that span the Hilbert space of $`N`$-particles $`|\psi >={\displaystyle \underset{\{a_j\},\{x_j\}}{}}\psi _{a_1,\mathrm{}a_N}(x_1,\mathrm{},x_N)C_{x_1a_1}^+\mathrm{}C_{x_Na_N}^+|0>.`$ where $`x_j\{1,2,\mathrm{},L\}`$, $`a_j\{1,2,3,4\}`$ and $`j=1,2,\mathrm{},N`$. The eigenvalue problem $`H|\psi >=E|\psi >`$ becomes an $`N`$-particle quantum mechanical problem with the first quantized Schrödinger operator (Hamiltonian), $$=t\underset{j=1}{\overset{N}{}}\mathrm{\Delta }_j+U\underset{i<j}{}(1\delta _{a_i,a_j})\delta (x_i,x_j),$$ (12) if site occupations of more than two electrons are excluded , where $`\mathrm{\Delta }_j\psi :=\psi (\mathrm{},x_j+1,\mathrm{})+\psi (\mathrm{},x_j1,\mathrm{})`$. The wave function of Bethe-ansatz form in the region $`x𝒞(Q):=\{x|\mathrm{\hspace{0.17em}1}x_{Q1}<\mathrm{}<x_{QN}L\}`$ reads $$\psi _a(x)=\underset{PS_N}{}A_a(P,Q)e^{i(Pk|x)},$$ (13) where $`x=(x_1,x_2,\mathrm{},x_N)`$, $`a:=(a_1`$, $`a_2,\mathrm{},a_N)`$, $`a_j`$ stands for the spin-orbital component of the “$`j`$th” particle; $`Pk`$ is the image of a given $`k:=(k_1,k_2,\mathrm{},k_N)`$ by a mapping $`PS_N`$; $`S_N`$ denotes the permutation group of $`N`$ objects; $`(Pk|x)=_{j=1}^N(Pk)_j(x)_j`$. The coefficients $`A(P,Q)`$ are functionals on $`S_NS_N`$. We known that any permutations can always be expressed as the product of the neighboring interchanges $`\mathrm{\Pi }^j:(\mathrm{},z_j,z_{j+1},\mathrm{})(\mathrm{},z_{j+1},z_j,\mathrm{})`$. So the requirement of antisymmetry for fermionic permutation is $`(\mathrm{\Pi }^j\psi )_a(x)=\psi _a(x)`$, which gives $$A(P,\mathrm{\Pi }^jQ)=𝒫^{Qj,Q(j+1)}A(\mathrm{\Pi }^jP,Q),$$ (14) where the spin-orbital labels are omitted and $`𝒫^{Qj,Q(j+1)}`$ is the SU(4) spinor representation of the permutation $`\mathrm{\Pi }^j`$ operator. An immediate consequence of (14) is $`\delta _{a_i,a_j}\delta (x_i,x_j)\psi _a(x)=0`$, and hence we are allowed to consider, instead of (12), the following equivalent Schrödinger operator $$=t\underset{j=1}{\overset{N}{}}\mathrm{\Delta }_j+U\underset{i<j}{}\delta (x_i,x_j),$$ (15) here the interaction terms are independent of the spin-orbital labels. Then the strategy of ref. can be used to solve the wave functions. The S-matrix that relates the coefficients $`A^{}s`$ between distinct regions in the configuration space of $`N`$ electrons can be solved from the Schrödinger equation in the vicinity of hyperplane with $`(Qx)_j=(Qx)_{j+1}`$. Accordingly, we get $$S^{Qj,Q(j+1)}=\frac{\mathrm{sin}(Pk)_{Qj}\mathrm{sin}(Pk)_{Q(j+1)}+ic𝒫^{Qj,Q(j+1)}}{\mathrm{sin}(Pk)_{Qj}\mathrm{sin}(Pk)_{Q(j+1)}+ic},$$ (16) where $`2c=U/t`$. As it satisfies Yang-Baxter equation , the Bethe-ansatz wave function is then consistently determined, i.e., the coefficients $`A`$’s in any region are determined up to an overall factor by the $`\stackrel{ˇ}{S}^{Qj,Q(j+1)}:=𝒫^{Qj,Q(j+1)}S^{Qj,Q(j+1)}`$ and that in different regions are related by (14). If let $`c\mathrm{}`$ in (16), we know the wave function will be null if there are two $`k`$’s being the same value. So in the strong coupling limit, the $`k`$’s must take distinct values though there can be four states corresponding to the same $`k_j`$ in the absence of interaction. The periodic boundary condition is guaranteed provided that $`A(P,\gamma Q)e^{i(Pk)_1L}=A(P,Q)`$ in which $`\gamma =\mathrm{\Pi }^{N1}\mathrm{\Pi }^{N2}\mathrm{}\mathrm{\Pi }^2\mathrm{\Pi }^1`$. After applying the S-matrices successively, one obtain an eigenvalue equation in the SU(4) spinor space: $`S^{Q1,QN}S^{Q1,Q(N1)}\mathrm{}S^{Q1,Q2}A(P,Q)`$ (17) $`=e^{i(Pk)_1L}A(P,Q).`$ (18) The eigenvalue problem (18) can be diagonalized by means of quantum inverse scattering method by defining the transfer matrix as $`t(\alpha )=tr𝒯(\alpha )`$ where $`𝒯(\alpha )=T_{AN}(\alpha \alpha _N)\mathrm{}T_{A2}(\alpha \alpha _2)T_{A1}(\alpha \alpha _1)`$, $`T_{Aj}(\alpha ):=S^{Aj}(\alpha )End(V^AV^j)`$. $`End`$ means endomorphism. It can also be diagonalized by similar procedure as in ref. where the general case of continuous model with $`\delta `$-function interaction was solved. The Bethe-ansatz equations reads $`e^{ik_jL}={\displaystyle \underset{\alpha =1}{\overset{M}{}}}\mathrm{\Xi }_{1/2}(\mathrm{sin}k_j\lambda _\alpha ),`$ (19) $`1`$ $`=`$ $`{\displaystyle \underset{l=1}{\overset{N}{}}}\mathrm{\Xi }_{1/2}(\lambda _\alpha \mathrm{sin}k_l){\displaystyle \underset{\alpha ^{}=1}{\overset{M}{}}}\mathrm{\Xi }_1(\lambda _\alpha \lambda _\alpha ^{}){\displaystyle \underset{\beta =1}{\overset{M^{}}{}}}\mathrm{\Xi }_{1/2}(\lambda _\alpha \mu _\beta ),`$ (20) $`1`$ $`=`$ $`{\displaystyle \underset{\alpha =1}{\overset{M}{}}}\mathrm{\Xi }_{1/2}(\mu _\beta \lambda _\alpha ){\displaystyle \underset{\beta ^{}=1}{\overset{M^{}}{}}}\mathrm{\Xi }_1(\mu _\beta \mu _\beta ^{}){\displaystyle \underset{\gamma =1}{\overset{M^{\prime \prime }}{}}}\mathrm{\Xi }_{1/2}(\mu _\beta \nu _\gamma ),`$ (21) $`1`$ $`=`$ $`{\displaystyle \underset{\beta =1}{\overset{M^{}}{}}}\mathrm{\Xi }_{1/2}(\nu _\gamma \mu _\beta ){\displaystyle \underset{\gamma ^{}=1}{\overset{M^{\prime \prime }}{}}}\mathrm{\Xi }_1(\nu _\gamma \nu _\gamma ^{}),`$ (22) where $`\mathrm{\Xi }_n(x):=[x+inc]/[xinc]`$. Eq.(22) was given firstly in . Because a particular chemical potential was introduced in the Hamiltonian, the Bethe-ansatz equation derived in is the same as the SU(4) Heisenberg model . We write out the Bethe-ansatz equation in a form so that it is easy to be remembered by means of the “Dynken diagram” of $`A_3`$ Lie algebra Where the dark dot is added to represent the charge rapidity $`k_j`$ which also has an angle of $`120^o`$ relative to the first simple root $`r_1`$. The subscripts in $`\mathrm{\Xi }`$ in eq.(22) are then related to the covariant components of the simple roots when the simple roots are chosen as non-orthogonal basis, accordingly, $`r_1=(1/2,\mathrm{\hspace{0.17em}1},1/2,0)`$, $`r_2=(0,1/2,\mathrm{\hspace{0.17em}1},1/2)`$, $`r_3=(0,0,1/2,\mathrm{\hspace{0.17em}1})`$. The highest weight vector $`𝐰=(w_1,w_2,w_3)`$ labeling the representation of SU(4) carried out by the corresponding eigenstates is given by $`w_1`$ $`=`$ $`N2M+M^{},`$ (23) $`w_2`$ $`=`$ $`M2M^{}+M^{\prime \prime },`$ (24) $`w_3`$ $`=`$ $`M^{}2M^{\prime \prime }.`$ (25) A set of coupled transcendental equations are derived by taking the logarithm of (22), $`k_j{\displaystyle \frac{1}{L}}{\displaystyle \underset{\alpha =1}{\overset{M}{}}}\mathrm{\Theta }_{1/2}(\mathrm{sin}k_j\lambda _\alpha )`$ $`=`$ $`{\displaystyle \frac{2\pi }{L}}h_j,`$ (26) $`{\displaystyle \underset{l=1}{\overset{N}{}}}\mathrm{\Theta }_{1/2}(\lambda _\alpha \mathrm{sin}k_l)+{\displaystyle \underset{\alpha ^{}=1}{\overset{M}{}}}\mathrm{\Theta }_1(\lambda _\alpha \lambda _\alpha ^{})+{\displaystyle \underset{\beta =1}{\overset{M^{}}{}}}\mathrm{\Theta }_{1/2}(\lambda _\alpha \mu _\beta )`$ $`=`$ $`2\pi I_\alpha ,`$ (27) $`{\displaystyle \underset{\alpha =1}{\overset{M}{}}}\mathrm{\Theta }_{1/2}(\mu _\beta \lambda _\alpha )+{\displaystyle \underset{\beta ^{}=1}{\overset{M^{}}{}}}\mathrm{\Theta }_1(\mu _\beta \mu _\beta ^{})+{\displaystyle \underset{\gamma =1}{\overset{M^{\prime \prime }}{}}}\mathrm{\Theta }_{1/2}(\mu _\beta \nu _\gamma )`$ $`=`$ $`2\pi J_\beta ,`$ (28) $`{\displaystyle \underset{\beta =1}{\overset{M^{}}{}}}\mathrm{\Theta }_{1/2}(\nu _\gamma \mu _\beta )+{\displaystyle \underset{\gamma ^{}=1}{\overset{M^{\prime \prime }}{}}}\mathrm{\Theta }_1(\nu _\gamma \nu _\gamma ^{})`$ $`=`$ $`2\pi K_\gamma ,`$ (29) where $`\mathrm{\Theta }_n(x):=2\mathrm{tan}^1(\frac{x}{nc})`$. The quantum number $`h_j`$ for charge rapidity $`k_j`$ takes integer or half-integer value depending on whether $`M1`$ is odd or even. The quantum numbers $`I_\alpha `$, $`J_\beta `$ or $`K_\gamma `$ for flavor (we refer for the spin-orbital double) rapidities $`\lambda _\alpha `$, $`\mu _\beta `$ or $`\nu _\gamma `$, take integer or half-integer values respectively depending on whether $`NMM^{}`$, $`MM^{}M^{\prime \prime }`$ or $`M^{}M^{\prime \prime }`$ is odd or even. These properties arise from the logarithm function. Once the roots are solved from the above equation (29), the energy and momentum will be calculated by $`E`$ $`=`$ $`2t{\displaystyle \underset{j=1}{\overset{N}{}}}\mathrm{cos}k_j,`$ (30) $`P`$ $`=`$ $`{\displaystyle \frac{2\pi }{L}}\left[{\displaystyle \underset{l=1}{\overset{N}{}}}h_l+{\displaystyle \underset{\alpha =1}{\overset{M}{}}}I_\alpha +{\displaystyle \underset{\beta =1}{\overset{M^{}}{}}}J_\beta +{\displaystyle \underset{\gamma =1}{\overset{M^{\prime \prime }}{}}}K_\gamma \right].`$ (31) ## III Ground state The ground state is nondegenerate only if $`N=4n`$ for $`n`$ being odd numbers. This is easily seen by considering non-interaction case. The momentum eigenvalues (with periodic boundary condition) of non-interacting electrons are $`k=m(2\pi /L)`$, $`m=0,\pm 1,\pm 2\mathrm{}`$. For example, $`N=4n`$ for $`n=even`$, the ground state has a 70-fold degeneracy. In the following we will restrict ourselves to the case of $`N_0=4n`$ for $`n=odd`$, and consider the non-degenerated ground state. The non-degenerated ground state is a SU(4) singlet which is characterized by a 4-row and n-column Young tableau. The quantum numbers in (29) for the ground state are $`\{h_j\}`$ $`=`$ $`\{(N_01)/2,\mathrm{},(N_01)/2\},`$ (32) $`\{I_\alpha \}`$ $`=`$ $`\{(3N_04)/8,\mathrm{},(3N_04)/8\},`$ (33) $`\{J_\beta \}`$ $`=`$ $`\{(N_02)/4,\mathrm{},(N_02)/4\},`$ (34) $`\{K_\gamma \}`$ $`=`$ $`\{(N_04)/8,\mathrm{},(N_04)/8\}.`$ (35) Obviously, $`h`$’s and $`J`$’s are consecutive half-integers while $`I`$’s and $`K`$’s are consecutive integers. As a result of (35) and (31), the momentum of the non-degenerate ground state is zero. We plot the ground-state energy with respect to the filling factor $`N/L`$ for various coupling constant in fig.(1). The relation between Fermi momentum and the filling factor for different coupling constant is also plotted there. These numerical results are calculated for $`L=20`$ and $`N`$ from $`4`$ to $`40`$. For $`N\mathrm{}`$, we can take thermodynamics limit which will be discussed in Sec.VI. The ground state for $`N=N_0+1`$, $`N=N_0+2`$ etc. are degenerate about which we will discuss later. ## IV Excited states ### A Excitations above the nondegenerate ground state The excited states are obtained by variation in the sequence of quantum numbers $`\{h_j\}`$, $`\{I_\alpha \}`$, $`\{J_\beta \}`$ or $`\{K_\gamma \}`$ from that for the ground state. The simplest case is to remove one of the $`h`$’s from the sequence of ground state (35) and add a new $`h_0`$ outside of the original sequence. That is $`\{h_j\}=\{{\displaystyle \frac{N_01}{2}},\mathrm{}{\displaystyle \frac{N_01}{2}}+n_01,`$ (36) $`{\displaystyle \frac{N_01}{2}}+n_0+1,\mathrm{},{\displaystyle \frac{N_01}{2}},h_0\},`$ (37) with $`|h_0|>(N_01)/2`$ and the other sequences in (35) keep unchanged. Clearly, the $`(N_01)/2+n_0`$ is absent in the set (37). In fig.(2), we plot the numerical results for energy-momentum spectrum, which is a two-parameter family of excitation. This kind of excitations are singlet states. There are several further possibilities. After moving one box from the fourth row in the Young tableau of the ground state to the first row, we get a Young tableau labeling a 15-dimensional irreducible representation of SU(4) according to the knowledge of group theory. It requires $`M=3N_0/41`$, $`M^{}=N_0/21`$ and $`M^{\prime \prime }=N_0/41`$. This causes the $`h`$’s and $`J`$’s to take integer values instead of half-integer values that were taken for ground state. There are now $`M`$ allowed values for the $`M1`$ distinct $`I`$’s, and $`M^{\prime \prime }`$ allowed values for the $`M^{\prime \prime }1`$ distinct $`K`$’s. Consequently, holes in the $`I`$’s and $`K`$’s sequences indispensably occurred, and then the low-energy states are parameterized by, $`\{h_j\}`$ $`=`$ $`\{N_0/2+1,\mathrm{},N_0/21,N_0/2\},`$ (38) $`\{J_\beta \}`$ $`=`$ $`\{N_0/4+1,\mathrm{},N_0/41\},`$ (39) $`I_1`$ $`=`$ $`{\displaystyle \frac{3N_04}{8}}+\delta _{1,\alpha _1},`$ (40) $`I_\alpha `$ $`=`$ $`I_{\alpha 1}+1+\delta _{\alpha ,\alpha _1}(\alpha =2,\mathrm{}{\displaystyle \frac{3N_0}{4}}1),`$ (41) $`K_1`$ $`=`$ $`{\displaystyle \frac{N_04}{8}}+\delta _{1,\gamma _1},`$ (42) $`K_\gamma `$ $`=`$ $`K_{\gamma 1}+1+\delta _{\gamma ,\gamma _1}(\gamma =2,\mathrm{},{\displaystyle \frac{N_0}{4}}1),`$ (43) where $`1\alpha _13N_0/4`$, $`1\gamma _1N_0/4`$. Numerical results of the energy-momentum spectra for this type of excitation are plotted in fig.(3), which is a two-parameter family of excitation. The states with negative momentum are just obtained by shifting the $`\{h_j\}`$ in (43) to the left by one unit. Moving one box from the fourth and one from the third row of the Young tableau for the ground state to the first and second row, we get the 20-fold excitation with $`M=3N_0/41`$, $`M^{}=N_0/22`$ and $`M^{\prime \prime }=N_0/41`$. The $`I`$’s and $`K`$’s that took integer values in the ground state now take half-integer values; the $`h`$’s that took half-integer values now takes integer values and the $`J`$’s still take half-integer values. As a result, two holes in $`\{J_\beta \}`$ are indispensably appeared, $`\{h_j\}`$ $`=`$ $`\{N_0/2+1,\mathrm{},N_0/21,N_0/2\},`$ (44) $`\{I_\alpha \}`$ $`=`$ $`\{3N_0/8+1,\mathrm{},3N_0/81\},`$ (45) $`\{K_\gamma \}`$ $`=`$ $`\{N_0/8+1,\mathrm{},N_0/81\},`$ (46) $`J_1`$ $`=`$ $`{\displaystyle \frac{N_02}{4}}+\delta _{1,\beta _1},`$ (47) $`J_\beta `$ $`=`$ $`J_{\beta 1}+1+\delta _{\beta ,\beta _1}+\delta _{\beta ,\beta _2},`$ (48) where $`\beta =2,\mathrm{},N_0/22`$ and $`1\beta _1\beta _2N_0/2`$. Numerical results of the energy-momentum spectra for this type of excitation, a two-parameter family, are given in fig.(4). Similarly, the states with negative momentum are just obtained by shifting the $`\{h_j\}`$ in (48) to the left by one unit. Moving one box from the fourth and one from the third row of the Young tableau for the ground state to the first row, we get the 45-fold excitation with $`M=3N_0/42`$, $`M^{}=N_0/22`$ and $`M^{\prime \prime }=N_0/41`$. This makes the $`J`$’s to take integer values instead of half-integer ones and $`K`$’s to take half-integer values instead of integer ones. In this case, there exist two holes in $`\{I_\alpha \}`$ and one hole in $`\{J_\beta \}`$, namely $`\{h_j\}`$ $`=`$ $`\{N_0/2+1,\mathrm{},N_0/21\},`$ (49) $`\{K_\gamma \}`$ $`=`$ $`\{N_0/8+1,\mathrm{},N_0/81\},`$ (50) $`I_1`$ $`=`$ $`{\displaystyle \frac{3N_04}{8}}+\delta _{1,\alpha _1},`$ (51) $`I_\alpha `$ $`=`$ $`I_{\alpha 1}+1+\delta _{\alpha ,\alpha _1}+\delta _{\alpha ,\alpha _2},`$ (52) $`J_1`$ $`=`$ $`{\displaystyle \frac{N_0}{4}}+1+\delta _{1,\beta _1},`$ (53) $`J_\beta `$ $`=`$ $`J_{\beta 1}+1+\delta _{\beta ,\beta _1}(\beta =2,\mathrm{},{\displaystyle \frac{N_0}{2}}2),`$ (54) where $`\alpha =2,\mathrm{},3N_0/42`$, $`1\alpha _1<\alpha _23N_0/4`$ and $`1\beta _1N_0/21`$. Numerical results are plotted in fig.(5). It is a three-parameter family of excitation. Taking out one box respectively from the second, the third and the fourth row and putting them together on the first row of the Young tableau, we have a 35-fold excitation. Due to $`M=3N_0/43`$, $`M^{}=N_0/22`$ and $`M^{\prime \prime }=N_0/41`$, we have four holes in $`\{I_\alpha \}`$, accordingly, $`\{h_j\}`$ $`=`$ $`\{N_0/2+1,\mathrm{},N_0/21,N_0/2\},`$ (55) $`\{J_\beta \}`$ $`=`$ $`\{(N_06)/4,\mathrm{},(N_06)/4\},`$ (56) $`\{K_\gamma \}`$ $`=`$ $`\{N_0/8+1,\mathrm{},N_0/81\},`$ (57) $`I_1`$ $`=`$ $`{\displaystyle \frac{3N_0}{4}}+\delta _{1,\alpha _1},`$ (58) $`I_\alpha `$ $`=`$ $`I_{\alpha 1}+1+{\displaystyle \underset{i=1}{\overset{4}{}}}\delta _{\alpha ,\alpha _i}(\alpha =2,\mathrm{},3N_0/43),`$ (59) where $`1\beta _1<\beta _2<\beta _3<\beta _43N_0/4+1`$. The numerical results are plotted in fig.(6), where we did not plot the pattern obtained by shifting $`\{h_j\}`$ in (59) to the left by one unit, which is just the mirror image of the plotted pattern. This is a four-parameter family of excitation. ### B Adding particles If the number of electrons are $`N_0+1`$, $`N_0+2`$ or $`N_0+3`$, the corresponding states can be obtained by adding one, two or three particles into the system of $`N_0`$ electrons. Adding one particle to the $`N_0`$ ground state and leaving $`M`$, $`M^{}`$ and $`M^{\prime \prime }`$ unchanged, the $`h`$’s, $`I`$’s and $`J`$’s are half-odd integers but $`K`$’s are integers. Comparing to that for the non-degenerate ground state of $`N_0=4n`$, there are now $`3N_0/4+1`$ allowed values for the $`3N_0/4`$ distinct $`I`$’s. So there is always a “hole” in the $`I`$’s sequence, namely, $`\{h_j\}`$ $`=`$ $`\{(N_01)/2,\mathrm{},(N_01)/2,h_0\},`$ (60) $`I_1`$ $`=`$ $`3N_0/8+\delta _{1,\alpha _1},`$ (61) $`I_\alpha `$ $`=`$ $`I_{\alpha 1}+1+\delta _{\alpha ,\alpha _1}(\alpha =2,\mathrm{},3N_0/4),`$ (62) where $`1\alpha _13N_0/4+1`$. The $`J`$’s and $`K`$’s sequences are the same as those in (35). The numerical results for $`h_0>0`$ are plotted in fig.(7), where the zero energy corresponds to the ground state for $`N=N_0+1`$. The spectra with negative momentum are obtained by using $`h_0<0`$. It is easy to know by evaluating eq.(25) that each point in the figure represents a quadruplet. Adding two particles to the $`N_0`$ particle ground state, we have $`N=N_0+2`$, $`M=3N_0/4+1`$, $`M^{}=N_0/2`$ and $`M^{\prime \prime }=N_0/4`$ for the low-energy states. This requires $`h`$’s, $`J`$’s and $`K`$’s to take integer values but $`I`$’s to take half of odd integer values. Referring to (35) we know that there must be a “hole” in the $`J`$’s sequence, consequently, $`\{h_j\}`$ $`=`$ $`\{N_0/2,\mathrm{},N_0/2,h_0\},`$ (63) $`\{I_\alpha \}`$ $`=`$ $`\{3N_0/8,\mathrm{},3N_0/8\},`$ (64) $`J_1`$ $`=`$ $`{\displaystyle \frac{N_0}{4}}+\delta _{1,\beta _1},`$ (65) $`J_\beta `$ $`=`$ $`J_{\beta 1}+1+\delta _{\beta ,\beta _1}(\beta =2,\mathrm{},N_0/2),`$ (66) where $`1\beta _1N_0/2+1`$. The $`K`$’s sequence is the same as that in (35). The numerical results are plotted in fig.(8). Adding three particles to the $`N_0`$ ground state, we have $`N=N_0+3`$, $`M=3N_0/4+2`$, $`M^{}=N_0/2+1`$ and $`M^{\prime \prime }=N_0/4`$ for low-energy states. This requires $`h`$’s and $`K`$’s to be half-odd integers but $`I`$’s and $`J`$’s to be integers, and then $`\{h_j\}`$ $`=`$ $`\{(N_0+1)/2,\mathrm{},(N_0+1)/2,h_0\},`$ (67) $`\{I_\alpha \}`$ $`=`$ $`\{(3N_0+4)/8,\mathrm{},(3N_0+4)/8\},`$ (68) $`\{J_\beta \}`$ $`=`$ $`\{N_0/8,\mathrm{},N_0/4\},`$ (69) $`K_1`$ $`=`$ $`{\displaystyle \frac{N_0}{8}}+\delta _{1,\gamma _1},`$ (70) $`K_\gamma `$ $`=`$ $`K_{\gamma 1}+1+\delta _{\gamma ,\gamma _1}(\gamma =2,\mathrm{},N_0/4),`$ (71) where $`1\gamma _1N_0/4+1`$. The numerical results are plotted in fig.(9). ## V Special cases For some special limiting cases, one is able to obtain several interesting conclusions from the Bethe-ansatz equation without solving it directly. In the following, we consider the weak coupling and strong coupling respectively. ### A Weak coupling Because $`\mathrm{\Theta }_n(x)\pi \mathrm{sign}(x)`$ for $`c0`$, the Bethe-ansatz equation (22) becomes $`k_j+{\displaystyle \frac{\pi }{L}}{\displaystyle \underset{\alpha =1}{\overset{M}{}}}\mathrm{sgn}(\mathrm{sin}k_j\lambda _\alpha )`$ $`=`$ $`{\displaystyle \frac{2\pi }{L}}h_j,`$ (72) $`{\displaystyle \underset{l=1}{\overset{N}{}}}\mathrm{sgn}(\lambda _\alpha \mathrm{sin}k_l){\displaystyle \underset{\alpha ^{}=1}{\overset{M}{}}}\mathrm{sgn}(\lambda _\alpha \lambda _\alpha ^{})+{\displaystyle \underset{\beta =1}{\overset{M^{}}{}}}\mathrm{sgn}(\lambda _\alpha \mu _\beta )`$ $`=`$ $`2I_\alpha ,`$ (73) $`{\displaystyle \underset{\alpha =1}{\overset{M}{}}}\mathrm{sgn}(\mu _\beta \lambda _\alpha ){\displaystyle \underset{\beta ^{}=1}{\overset{M^{}}{}}}\mathrm{sgn}(\mu _\beta \mu _\beta ^{})+{\displaystyle \underset{\gamma =1}{\overset{M^{\prime \prime }}{}}}\mathrm{sgn}(\mu _\beta \nu _\gamma )`$ $`=`$ $`2J_\beta ,`$ (74) $`{\displaystyle \underset{\beta =1}{\overset{M^{}}{}}}\mathrm{sgn}(\nu _\gamma \mu _\beta ){\displaystyle \underset{\gamma ^{}=1}{\overset{M^{\prime \prime }}{}}}\mathrm{sgn}(\nu _\gamma \nu _\gamma ^{})`$ $`=`$ $`2K_\gamma .`$ (75) Without loss of generality, the $`\gamma `$ label can be so chosen that $`K_\gamma `$ is arranged in an increasing order. Then the fourth equation of (75) becomes $`{\displaystyle \underset{\mu =1}{\overset{M^{}}{}}}\mathrm{sgn}(\nu _\gamma \mu _\beta )=2K_\gamma +2\gamma M^{\prime \prime }1.`$ (76) Because $`|K_\gamma |(M^{\prime \prime }1)/2`$ and $`M^{\prime \prime }M^{}/2`$ (restricted by the Young tableau), the minimum value of right-hand side of (76) is $`M^{}+2`$. This means that the smallest $`\mu _\beta `$ is smaller than the smallest $`\nu _\gamma `$. From eq.(76) we easily derive $`{\displaystyle \underset{\beta =1}{\overset{M^{}}{}}}[\mathrm{sgn}(\nu _{\gamma +1}\mu _\beta )\mathrm{sgn}(\nu _\gamma \mu _\beta )]`$ (77) $`=2(K_{\gamma +1}K_\gamma +1).`$ (78) Obviously, if $`K_{\gamma +1}K_\gamma =m^{\prime \prime }`$, there must exist exactly $`m^{\prime \prime }+1`$ solutions $`\mu _\beta `$ satisfying $`\nu _\gamma <\mu _\beta <\nu _{\gamma +1}`$. Again from the third equation of (75), we have $`{\displaystyle \underset{\alpha =1}{\overset{M}{}}}[\mathrm{sgn}(\mu _{\beta +1}\lambda _\alpha )\mathrm{sgn}(\mu _\beta \lambda _\alpha )]=2J_{\beta +1}2J_\beta `$ (79) $`+2{\displaystyle \underset{\gamma =1}{\overset{M^{\prime \prime }}{}}}[\mathrm{sgn}(\mu _{\beta +1}\nu _\gamma )\mathrm{sgn}(\mu _\beta \nu _\gamma )].`$ (80) Obviously, if there is a $`\nu _\gamma `$ such that $`\mu _\beta <\nu _\gamma <\mu _{\beta +1}`$ then the right hand side of (80) equals $`2(J_{\beta +1}J_\beta )`$, otherwise it equals $`2(J_{\beta +1}J_\beta +1)`$. Proceeding the same strategy to the second equation and the first equation in (75) successively, we obtain a sequence of relations that are summarized as follows. (i) If $`K_{\gamma +1}K_\gamma =m^{\prime \prime }`$, there exist exactly $`m^{\prime \prime }+1`$ solutions $`\mu _\beta `$ satisfying $`\nu _\gamma <\mu _\beta <\nu _{\gamma +1}`$. (ii) For $`J_{\beta +1}J_\beta =m^{}`$, if there exists a $`\nu _\gamma `$ satisfying $`\mu _\beta <\nu _\gamma <\mu _{\beta +1}`$, there will be $`m^{}`$ $`\lambda `$’s such that $`\mu _\beta <\lambda _\alpha <\mu _{\beta +1}`$; otherwise if there is no $`\nu _\gamma `$ satisfying that, then there will be $`m^{}+1`$ $`\lambda `$’s such that $`\mu _\beta <\lambda _\alpha <\mu _{\beta +1}`$. (iii) For $`I_{\alpha +1}J_\alpha =m`$, if there is a $`\mu _\beta `$ satisfying $`\lambda _\alpha <\mu _\beta <\lambda _{\alpha +1}`$, there will exist $`m`$ $`\mathrm{sin}k_l`$’s such that $`\lambda _\alpha <\mathrm{sin}k_l<\lambda _{\alpha +1}`$; otherwise there will be $`m+1`$ $`\mathrm{sin}k_l`$’s such that $`\lambda _\alpha <\mathrm{sin}k_l<\lambda _{\alpha +1}`$ (iv) For $`h_{j+1}h_j=n^{}`$, if there exists such a $`\lambda _\alpha `$ that $`\mathrm{sin}k_j<\lambda _\alpha <\mathrm{sin}k_{j+1}`$, then we will have $`k_{j+1}k_j=2\pi (n^{}1)/L`$; otherwise we will have $`k_{j+1}k_j=2\pi n^{}/L`$. Apply these items to the ground state $`n^{}=m=m^{}=m^{\prime \prime }=1`$, we conclude that the sequence $`\{k_j\}`$ is divided into groups with four successive values in each group. Between each pair $`(k_{4j3},k_{4j2})`$, $`(k_{4j2},k_{4j1})`$ or $`(k_{4j1},k_{4j})`$ in the same group there is one $`\lambda `$, so totally there are three $`\lambda `$’s in each group. Furthermore there are two $`\mu `$’s each lying in between the adjacent $`\lambda `$’s. And between these two $`\mu `$’s there is always a $`\nu `$. ### B Strong coupling For the strong coupling limit case, $`c\mathrm{}`$ (i.e. large $`U`$ limit), the ration $`\mathrm{sin}k_j/c`$ in the Bethe-ansatz equation is neglectable. The first equation of (22) then becomes $`k_j`$ $`=`$ $`{\displaystyle \frac{2\pi }{L}}n_j{\displaystyle \frac{1}{L}}P_H,`$ (81) $`P_H`$ $`=`$ $`{\displaystyle \underset{\alpha =1}{\overset{M}{}}}\left[\pi 2\mathrm{tan}^1(2\lambda _\alpha /c)\right],`$ (82) where $`n_j`$ are integers. Obviously, the $`P_H`$ is just the momentum of spin-orbital excitations in the SU(4) Heisenberg chain. The other three Bethe-ansatz equations turned to be those known for the SU(4) Heisenberg chain for the scaled rapidities, $`\lambda _\alpha /c`$, $`\mu _\beta /c`$ and $`\nu _\gamma /c`$. Equation (82) indicates that the allowed quasi-momenta $`k`$’s in the strong coupling limit are quantized in units of $`2\pi /L`$, just like “spinless” noninteracting fermions. This is because the double occupancy is forbidden by the strong repulsive on-site coupling. The allowed quasimomenta due to periodic boundary condition are determined by spin-orbital momentum $`P_H`$. Particularly, for $`N=4n`$ with $`n=odd`$, the ground-state spin-orbital momentum $`P_H`$ is an odd multiple of $`\pi `$ so that the allowed $`k`$’s are half odd integers multiplied by $`2\pi /L`$. Therefore, the ground state for $`N=4n`$ ($`n=odd`$) is uniquely determined, i.e., it is nondegenerate. This is different from the noninteracting case in which the allowed $`k`$’s are always integers multiplied by $`2\pi /L`$ for any $`N`$ and for each $`k`$ there are four states because of the spin-orbital degree of freedom. ## VI Thermodynamics limit Replacing $`k_j`$, $`\lambda _\alpha `$, $`\mu _\beta `$ and $`\nu _\gamma `$ in eq. (29) by continuous variables $`k`$, $`\lambda `$, $`\mu `$, and $`\nu `$ but keeping the summation still over the solution set of these roots, we can consider the quantum numbers $`h_j`$, $`I_\alpha `$, $`J_\beta `$, and $`K_\gamma `$ as functions $`h(k)`$, $`I(\lambda )`$, $`J(\mu )`$, and $`K(\nu )`$ given by eq. (29). Take $`I(\lambda )`$ as an example, when $`I(\lambda )`$ passes through one of the quantum numbers $`I_j`$, the corresponding $`\lambda `$ is equal to one of the roots $`\lambda _j`$, similarly for $`J(\mu )`$, $`K(\nu )`$, or $`h(k)`$. However, there may exist some integers or half-integers for which the corresponding $`\lambda `$ ($`\mu `$, $`\nu `$, or $`k`$) is not in the set of roots. Such a situation is conventionally referred as a “hole”. In the thermodynamics limit, $`N\mathrm{}`$, $`L\mathrm{}`$, but $`N/L`$ kept finite, we may introduce the density of real roots and the density of holes (indicated by a subscript $`h`$), $`\rho (k)+\rho _h(k)=(1/L)dh(k)/dk,`$ (83) $`\sigma (\lambda )+\sigma _h(\lambda )=(1/L)dI(\lambda )/d\lambda ,`$ (84) $`\omega (\mu )+\omega _h(\mu )=(1/L)dJ(\mu )/d\mu ,`$ (85) $`\tau (\nu )+\tau _h(\nu )=(1/L)dK(\nu )/d\nu .`$ (86) By replacing the summations by integrals, for example, $`\underset{L\mathrm{}}{lim}{\displaystyle \frac{1}{L}}{\displaystyle \underset{l=1}{\overset{N}{}}}f(k_l)={\displaystyle _Q^Q}𝑑k\rho (k)f(k),`$ (87) $`\underset{L\mathrm{}}{lim}{\displaystyle \frac{1}{L}}{\displaystyle \underset{\alpha =1}{\overset{M}{}}}g(\lambda _\alpha )={\displaystyle _B^B}𝑑\lambda \sigma (\lambda )g(\lambda ),`$ (88) and so forth, eq.(29) gives rise to the following coupled integral equations, $`\rho (k)+\rho ^{(o)}(k)`$ $`=`$ $`{\displaystyle \frac{1}{2\pi }}\mathrm{cos}k{\displaystyle _B^B}K_{1/2}(\mathrm{sin}k\lambda )\sigma (\lambda )𝑑\lambda ,`$ (89) $`\sigma (\lambda )+\sigma ^{(o)}(\lambda )`$ $`=`$ $`{\displaystyle _Q^Q}K_{1/2}(\lambda \mathrm{sin}k)\rho (k)𝑑k{\displaystyle _B^B}K_1(\lambda \lambda ^{})\sigma (\lambda ^{})𝑑\lambda ^{}{\displaystyle _B^{}^B^{}}K_{1/2}(\lambda \mu )\omega (\mu )𝑑\mu ,`$ (90) $`\omega (\mu )+\omega ^{(o)}(\mu )`$ $`=`$ $`{\displaystyle _B^B}K_{1/2}(\mu \lambda )\sigma (\lambda )𝑑\lambda {\displaystyle _B^{}^B^{}}K_1(\mu \mu ^{})\omega (\mu ^{})𝑑\mu ^{}{\displaystyle _{B^{\prime \prime }}^{B^{\prime \prime }}}K_{1/2}(\mu \nu )\tau (\nu )𝑑\nu ,`$ (91) $`\tau (\nu )+\tau ^{(o)}(\nu )`$ $`=`$ $`{\displaystyle _B^{}^B^{}}K_{1/2}(\nu \mu )\omega (\mu )𝑑\mu {\displaystyle _{B^{\prime \prime }}^{B^{\prime \prime }}}K_1(\nu \nu ^{})\tau (\nu ^{})𝑑\nu ^{},`$ (92) where $`K_n(x):=\pi ^1nc/(n^2c^2+x^2)`$, and $`Q`$, $`B`$, $`B^{}`$, and $`B^{\prime \prime }`$ in the definite integrals should be determined self-consistently by $`{\displaystyle \frac{N}{L}}`$ $`=`$ $`{\displaystyle _Q^Q}\rho (k)𝑑k,`$ (93) $`{\displaystyle \frac{M}{L}}`$ $`=`$ $`{\displaystyle _B^B}\sigma (\lambda )𝑑\lambda ,`$ (94) $`{\displaystyle \frac{M^{}}{L}}`$ $`=`$ $`{\displaystyle _B^{}^B^{}}\omega (\mu )𝑑\mu ,`$ (95) $`{\displaystyle \frac{M^{\prime \prime }}{L}}`$ $`=`$ $`{\displaystyle _{B^{\prime \prime }}^{B^{\prime \prime }}}\tau (\nu )𝑑\nu .`$ (96) They hold for the case in the absence of the complex roots. In the presence of complex roots, however, it has variants. In eq.(92) we denoted the inhomogeneous terms by $`\rho ^{(o)}`$, $`\sigma ^{(o)}`$, $`\omega ^{(o)}`$ and $`\tau ^{(o)}`$, which not only stand for the densities of holes $`\rho _h`$, $`\sigma _h`$ etc., but also the contributions from complex roots, two-strings. Once the density $`\rho (k)`$ is solved from eq.(92), we have the $`z`$-components of the total spin and the total orbital $`{\displaystyle \frac{S_{tot}^z}{L}}`$ $`=`$ $`{\displaystyle \frac{1}{2}}{\displaystyle _Q^Q}\rho (k)𝑑k+{\displaystyle _B^{}^B^{}}\omega (\mu )𝑑\mu `$ (97) $`{\displaystyle _B^B}\sigma (\lambda )𝑑\lambda {\displaystyle _{B^{\prime \prime }}^{B^{\prime \prime }}}\tau (\nu )𝑑\nu ,`$ (98) $`{\displaystyle \frac{T_{tot}^z}{L}}`$ $`=`$ $`{\displaystyle \frac{1}{2}}{\displaystyle _Q^Q}\rho (k)𝑑k{\displaystyle _B^{}^B^{}}\omega (\mu )𝑑\mu .`$ (99) This is useful for a correct calculation of magnetizations. The energy is given by $$\frac{E}{L}=2t_Q^Q\mathrm{cos}k\rho (k)𝑑k.$$ (100) The highest weight vector that characterizes the corresponding representation is given by $`w_1`$ $`=`$ $`L{\displaystyle _Q^Q}\rho (k)𝑑k2L{\displaystyle _B^B}\sigma (\lambda )𝑑\lambda +L{\displaystyle _B^{}^B^{}}\omega (\mu )𝑑\mu ,`$ (101) $`w_2`$ $`=`$ $`L{\displaystyle _B^B}\sigma (\lambda )𝑑\lambda 2L{\displaystyle _B^{}^B^{}}\omega (\mu )𝑑\mu +L{\displaystyle _{B^{\prime \prime }}^{B^{\prime \prime }}}\tau (\nu )𝑑\nu ,`$ (102) $`w_3`$ $`=`$ $`L{\displaystyle _B^{}^B^{}}\omega (\mu )𝑑\mu 2L{\displaystyle _{B^{\prime \prime }}^{B^{\prime \prime }}}\tau (\nu )𝑑\nu .`$ (103) ### A Ground state properties The ground state of the present model is a Fermi sea described by $`\rho _0(k)`$, which is the distribution function of charge with respect to momentum $`k`$. The $`\tau _0(\nu )`$ describes the distribution of states with spin down and orbital bottom in the $`\nu `$-rapidity space. The $`\omega _0(\mu )`$ represents the distribution of either the state with spin up while orbital bottom or that with spin down while orbital bottom in the $`\mu `$-rapidity space. The $`\sigma _0(\lambda )`$, however, stands for the distribution of either state $`|2`$, $`|3`$ or $`|4`$ in the $`\lambda `$-rapidity space. These distribution functions satisfy (92) with $`B=B^{}=B^{\prime \prime }=\mathrm{}`$ and no holes, i.e., $`\rho ^{(o)}=0`$, $`\sigma ^{(o)}=0`$, $`\omega ^{(o)}=0`$, $`\tau ^{(o)}=0`$. By making Fourier transform to the second till the fourth equation of eq.(92), we have $`\stackrel{~}{\sigma }(q)`$ $`=`$ $`{\displaystyle \frac{1}{\sqrt{2\pi }}}{\displaystyle _{k_F}^{k_F}}e^{c|q|/2+iq\mathrm{sin}k}\rho (k)𝑑k`$ (104) $`\stackrel{~}{\sigma }(q)e^{c|q|}+\stackrel{~}{\omega }(q)e^{c|q|/2},`$ (105) $`\stackrel{~}{\omega }(q)`$ $`=`$ $`\stackrel{~}{\sigma }(q)e^{c|q|/2}\stackrel{~}{\omega }(q)e^{c|q|}+\stackrel{~}{\tau }(q)e^{c|q|/2},`$ (106) $`\stackrel{~}{\tau }(q)`$ $`=`$ $`\stackrel{~}{\omega }(q)e^{c|q|/2}\stackrel{~}{\tau }(q)e^{c|q|}.`$ (107) It is not difficult to obtain a single integral equation that determines the $`\rho _0(k)`$ $$\rho _0(k)=\frac{1}{2\pi }+\frac{\mathrm{cos}k}{c}_{k_F}^{k_F}R_{3/2}\left(\frac{\mathrm{sin}ksink^{}}{c}\right)\rho _0(k^{})𝑑k^{},$$ (108) where $`k_F`$ is the Fermi momentum and $`R_n(x)={\displaystyle _{\mathrm{}}^{\mathrm{}}}{\displaystyle \frac{dq}{2\pi }}{\displaystyle \frac{\mathrm{sinh}(nq)}{\mathrm{sinh}(2q)}}e^{iqx|q|/2}.`$ Once $`\rho _0(k)`$ is solved, the energy will be evaluated by the integral $$E_0/L=2t_{k_F}^{k_F}\mathrm{cos}k\rho _0(k)𝑑k.$$ Though an explicit expression cannot be obtained from eq.(108) in the general case, it becomes easier for a numerical calculation. It is immediate from eq.(107) that the highest weight vector (103) of the ground state is a null vector. Therefore the ground state is a SU(4) singlet, accordingly, both spin and orbital are “anti-ferromagnetic”. ### B ground-state energy for strong coupling The ground-state energy can be calculated explicitly at strong on-site coupling, $`c1`$ (i.e., $`Ut`$). Because of $`(\mathrm{sin}k\mathrm{sin}k^{})/c1`$ in this case and $`4\pi R_{3/2}(0)=3\mathrm{ln}2+\pi /2`$, eq. (108) is written out up to the order $`O(1/c)`$, $`\rho _0^{stro}={\displaystyle \frac{1}{2\pi }}+(3\mathrm{ln}2+{\displaystyle \frac{\pi }{2}}){\displaystyle \frac{\mathrm{cos}k}{4\pi c}}{\displaystyle \frac{N}{L}}.`$ The Fermi momentum determined from $`N/L=_{k_F}^{k_F}\rho _0(k)𝑑k`$ is $`k_F={\displaystyle \frac{N}{L}}\left[\pi (3\mathrm{ln}2+{\displaystyle \frac{\pi }{2}}){\displaystyle \frac{\mathrm{sin}(\pi N/L)}{2c}}\right],`$ and then the energy is calculated $`{\displaystyle \frac{E_0}{L}}`$ $`=`$ $`{\displaystyle \frac{t\mathrm{sin}(\pi N/L)}{\pi /2}}`$ (109) $``$ $`{\displaystyle \frac{t^2}{U}}({\displaystyle \frac{N}{L}})^2(3\mathrm{ln}2+{\displaystyle \frac{\pi }{2}})\left[1{\displaystyle \frac{\mathrm{sin}(2\pi N/L)}{2\pi N/L}}\right],`$ (110) where $`N/L`$ is the filling factor. It becomes $`E_0^{(1/2)}=Nt^2(6\mathrm{ln}2+\pi )/U`$ at half-filling $`N=2L`$. At quarter filling $`L=N`$ we have $`{\displaystyle \frac{E_0^{(1/4)}}{N}}=2{\displaystyle \frac{t^2}{U}}({\displaystyle \frac{3}{2}}\mathrm{ln}2+{\displaystyle \frac{\pi }{4}}),`$ which agrees with the result of the SU(4) Heisenberg model for $`J=2t^2/U`$. Because the model is solved under the assumption of excluding site occupations of more than two, the results here are not valid for above half-filling $`N>2L`$ in which there must exist sites occupied by three electrons and the Bethe-ansatz wave-function failures at that point in configuration space. However, the energy is expected to be evaluate by a particle-hole transformation, $$E(N/L,U)/L=E(4N/L,U)/L+3U(N/L2).$$ (111) In the next section and thereafter we will study low-lying excitations on the basis of the thermodynamics limit. ## VII Spin-orbital excitations It is convenient to study the excitations by introducing $`\rho (k)=\rho _0(k)+\rho _1(k)/L`$, $`\sigma (\lambda )=\sigma _0(\lambda )+\sigma _1(\lambda )/L`$, $`\omega (\mu )=\omega _0(\mu )+\omega _1(\mu )/L`$, and $`\tau (\nu )=\tau _0(\nu )+\tau _1(\nu )/L`$ where $`\rho _0(k)`$, $`\sigma _0(\lambda )`$, $`\omega _0(\mu )`$, and $`\tau _0(\nu )`$ satisfy the same set of integral equations as the ground state did. The excitation energy up to the order $`O(1/L)`$ is $$\mathrm{\Delta }E=_Q^Q𝑑k(2t\mathrm{cos}k+\mathrm{\Lambda })\rho _1(k),$$ (112) where $`\mathrm{\Lambda }`$ stands for the chemical potential . $`Q`$ can be replaced by $`k_F`$ for a large system. Equation (112) is valid for both the spin-orbital excitation and the charge excitation. The excitation energy is related to $`\rho _1(k)`$, which, moreover, should be solved from the following coupled integral equations: $`\rho _1(k)+\rho _1^{(o)}(k)`$ $`=`$ $`\mathrm{cos}k{\displaystyle _{\mathrm{}}^{\mathrm{}}}K_{1/2}(\mathrm{sin}k\lambda )\sigma _1(\lambda )𝑑\lambda ,`$ (113) $`\sigma _1(\lambda )+\sigma _1^{(o)}(\lambda )`$ $`=`$ $`{\displaystyle _{k_F}^{k_F}}K_{1/2}(\lambda \mathrm{sin}k)\rho _1(k)𝑑k{\displaystyle _{\mathrm{}}^{\mathrm{}}}K_1(\lambda \lambda ^{})\sigma _1(\lambda ^{})𝑑\lambda ^{}{\displaystyle _{\mathrm{}}^{\mathrm{}}}K_{1/2}(\lambda \mu )\omega _1(\mu )𝑑\mu ,`$ (114) $`\omega _1(\mu )+\omega _1^{(o)}(\mu )`$ $`=`$ $`{\displaystyle _{\mathrm{}}^{\mathrm{}}}K_{1/2}(\mu \lambda )\sigma _1(\lambda )𝑑\lambda {\displaystyle _{\mathrm{}}^{\mathrm{}}}K_1(\mu \mu ^{})\omega _1(\mu ^{})𝑑\mu ^{}{\displaystyle _{\mathrm{}}^{\mathrm{}}}K_{1/2}(\mu \nu )\tau _1(\nu )𝑑\nu ,`$ (115) $`\tau _1(\nu )+\tau _1^{(o)}(\nu )`$ $`=`$ $`{\displaystyle _{\mathrm{}}^{\mathrm{}}}K_{1/2}(\nu \mu )\omega _1(\mu )𝑑\mu +{\displaystyle _{\mathrm{}}^{\mathrm{}}}K_1(\nu \nu ^{})\tau _1(\nu ^{})𝑑\nu ^{}.`$ (116) The limits for the definite integrals are the same as that for ground state, which are valid for the low-lying excitations. Beyond low-lying excitations, however, the integration limits $`Q`$, $`B`$, $`B^{}`$, and $`B^{\prime \prime }`$ should be determined consistently. Here we only consider low-lying excitations. In order to consider the excitations above the singlet ground state, we must analyze the decomposition of the direct product of the SU(4) fundamental representation for $`N=4n`$. Using the Young tableau, we can obtain that the decomposition gives rise to a direct sum of a series of irreducible representations, i.e., $`(0,0,0)`$, $`(1,0,1)`$, $`(0,2,0)`$, $`(2,1,0)`$, $`(4,0,0)`$ $`\mathrm{𝑒𝑡𝑐}.`$ So the excitation states in spin-orbital sector include both the singlet $`(0,0,0)`$ and the multiplets of 15-fold $`(1,0,1)`$, of 20-fold $`(0,2,0)`$, and of 45-fold $`(2,1,0)`$ or of 35-fold $`(4,0,0)`$ etc. After evaluating the contributions of roots and two-strings to the highest weight vectors that characterize the irreducible representations of SU(4), we can get the correct compositions of holes and two-strings that create the possible excitations allowed by group theory. ### A The multiplets One $`\lambda `$ hole and one $`\nu `$ hole together create a 15-fold multiplet. Let $`\sigma _1^{(o)}(\lambda )=\sigma ^h(\lambda )=\delta (\lambda \overline{\lambda })`$, $`\tau _1^{(o)}(\nu )=\tau ^h(\nu )=\delta (\nu \overline{\nu })`$, and the other inhomogeneous terms in eq.(116) be null. Equation (116) is reduced to a closed form by Fourier transform. The excitation energy is composed of two terms $$\mathrm{\Delta }E_{(15)}=\epsilon _\sigma (\overline{\lambda })+\epsilon _\tau (\overline{\nu }),$$ and each of them can be identified as a flavoron with energy $$\epsilon _f(\overline{x})=_{k_F}^{k_F}𝑑k(2t\mathrm{cos}k+\mathrm{\Lambda })\rho _1^f(k,\overline{x}),$$ (117) where $`f=\sigma `$, $`\tau `$, or $`\omega `$. The $`\rho _1^\sigma (k,\overline{\lambda })`$ is solved by $`\rho _1^\sigma (k,\overline{\lambda })+{\displaystyle \frac{\mathrm{cos}k/(4c)}{\sqrt{2}\mathrm{cosh}[{\displaystyle \frac{\pi }{2c}}(\mathrm{sin}k\overline{\lambda })]1}}=`$ (118) $`{\displaystyle \frac{\mathrm{cos}k}{c}}{\displaystyle _{k_F}^{k_F}}R_{3/2}\left({\displaystyle \frac{\mathrm{sin}ksink^{}}{c}}\right)\rho _1^\sigma (k^{},\overline{\lambda })𝑑k^{},`$ (119) and the $`\rho _1^\tau (k,\overline{\nu })`$ satisfies the following integral equation: $`\rho _1^\tau (k,\overline{\nu })+{\displaystyle \frac{\mathrm{cos}k/(4c)}{\sqrt{2}\mathrm{cosh}[{\displaystyle \frac{\pi }{2c}}(\mathrm{sin}k\overline{\nu })]+1}}=`$ (120) $`{\displaystyle \frac{\mathrm{cos}k}{c}}{\displaystyle _{k_F}^{k_F}}R_{3/2}\left({\displaystyle \frac{\mathrm{sin}ksink^{}}{c}}\right)\rho _1^\tau (k^{},\overline{\nu })𝑑k^{}.`$ (121) Two $`\mu `$ holes, $`\omega ^{(0)}(\mu )=\delta (\mu \overline{\mu }_1)+\delta (\mu \overline{\mu }_1)`$ create a 20-fold multiplet with excitation energy $$\mathrm{\Delta }E_{(20)}=\epsilon _\omega (\overline{\mu }_1)+\epsilon _\omega (\overline{\mu }_2),$$ where $`\epsilon _\omega (x)`$ is evaluated by the same integral (117), but $`\rho _1^\omega (k,\overline{\mu })`$ should solve $`\rho _1^\omega (k,\overline{\mu })+{\displaystyle \frac{\mathrm{cos}k/(4c)}{\mathrm{cosh}[{\displaystyle \frac{\pi }{2c}}(\mathrm{sin}k\overline{\mu })]}}=`$ (122) $`{\displaystyle \frac{\mathrm{cos}k}{c}}{\displaystyle _{k_F}^{k_F}}R_{3/2}\left({\displaystyle \frac{\mathrm{sin}ksink^{}}{c}}\right)\rho _1^\omega (k^{},\overline{\mu })𝑑k^{}.`$ (123) The 45-fold multiplet is a three-hole state created by two $`\lambda `$ holes and one $`\mu `$ hole, i.e., $`\sigma _1^{(o)}(\lambda )=\delta (\lambda \overline{\lambda }_1)+\delta (\lambda \overline{\lambda }_2)`$ and $`\omega _1^{(o)}(\mu )=\delta (\mu \overline{\mu })`$ for which the excitation energy is $$\mathrm{\Delta }E_{(45)}=\epsilon _\sigma (\overline{\lambda }_1)+\epsilon _\sigma (\overline{\lambda }_2)+\epsilon _\omega (\overline{\mu }).$$ Four $`\lambda `$ holes create a 35-fold multiplet with excitation energy $$\mathrm{\Delta }E_{(35)}=\underset{j=1}{\overset{4}{}}\epsilon _\sigma (\overline{\lambda }_j).$$ In the above we have seen that there are three types of elementary excitation modes in the spin-orbital sector(let us call the SU(4) flavor degree of freedom). We refer these three elementary excitation modes the flavorons. It is easy to know the contributions of the holes to the highest weight vectors, and to the spin and orbital. Consequently, the quadruplets $`(1,0,0)`$ or $`(0,0,1)`$ are flavorons carrying both spin $`1/2`$ and orbital $`1/2`$ with energies $`ϵ_\sigma `$ or $`ϵ_\tau `$, whereas the hexaplet $`(0,1,0)`$ is a flavoron carrying either spin $`1`$ or orbital $`1`$ with energy $`ϵ_\omega `$. Clearly, spins and orbitals are mixed up in present isotropic on-site coupling. The spin orbital separation is expected to occur for the anisotropic cases that can be caused by Hund’s rule. From eq.(119-123) we find that the asymptotic behavior of all the three densities of roots vanish as the rapidities go to infinity. Thus these elementary excitations are gapless, i.e. $`ϵ_f(\pm \mathrm{})=0`$. ### B The singlet By observing the contributions of two-strings to the highest weight vectors, we find that the flavorons can compound to form a singlet. In addition to placing to the $`\lambda `$-rapidity a hole at $`\overline{\lambda }`$ and to the $`\nu `$-rapidity a hole at $`\overline{\nu }`$, we take into account of three two-strings respectively in those three rapidities, $`\lambda ^\pm =\lambda _0\pm ic/2`$, $`\mu ^\pm =\mu _0\pm ic/2`$ and $`\nu ^\pm =\nu _0\pm ic/2`$. In this case, $`M/L`$ $`=`$ $`{\displaystyle _{\mathrm{}}^{\mathrm{}}}\sigma (\lambda )𝑑\lambda +2,`$ (124) $`M^{}/L`$ $`=`$ $`{\displaystyle _{\mathrm{}}^{\mathrm{}}}\omega (\mu )𝑑\mu +2,`$ (125) $`M^{\prime \prime }/L`$ $`=`$ $`{\displaystyle _{\mathrm{}}^{\mathrm{}}}\tau (\nu )𝑑\nu +2,`$ (126) and the inhomogeneous terms of eq.(116) read $`\rho _1^{(o)}(k)`$ $`=`$ $`\mathrm{cos}kK_1(\mathrm{sin}k\lambda _0),`$ (127) $`\sigma _1^{(o)}(\lambda )`$ $`=`$ $`\delta (\lambda \overline{\lambda })+K_{3/2}(\lambda \lambda _0)`$ (128) $`+K_{1/2}(\lambda _{}\lambda _0)K_1(\lambda \lambda _0),`$ (129) $`\omega _1^{(o)}(\mu )`$ $`=`$ $`K_{3/2}(\mu \mu _0)+K_{1/2}(\mu _{}\mu _0)`$ (130) $`K_1(\mu \lambda _0)K_1(\mu \nu _0),`$ (131) $`\tau _1^{(o)}(\nu )`$ $`=`$ $`\delta (\nu \nu _0)+K_{3/2}(\nu \nu _0)`$ (132) $`+K_{1/2}(\nu \nu _0)K_1(\nu \mu _0).`$ (133) Substituting them into eq.(116) and taking Fourier transform, we find that the term containing $`\nu _0`$ in the fourth equation cancels with the term containing $`\nu _0`$ in the third equation. After substituting the result into the second equation, again the terms containing $`\mu _0`$ cancel each other. The substituting of the obtained expression of $`\sigma _1(\lambda )`$ into the first equation brings about an exact cancellation of the terms containing $`\lambda _0`$. As a result, the excitation energy is obtained $$\mathrm{\Delta }E_{(1)}=\epsilon _\sigma (\overline{\lambda })+\epsilon _\tau (\overline{\nu }),$$ where $`\epsilon _\sigma `$ and $`\epsilon _\tau `$ are evaluated by eq.(117) in which the $`\rho _1^\sigma `$ and $`\rho _1^\tau `$ satisfy the eq.(119) and eq.(121), respectively. Clearly, the excitations of 15-fold multiplet and the singlet are degenerate in energy. ## VIII Charge excitations ### A The holon-antiholon excitation Let us consider the case of less than quarter-filling ($`N<L`$). we are allowed to add one “particle” outside the charge Fermi sea, $`k_p[k_F,k_F]`$ but leaving a hole inside the charge Fermi sea, $`\overline{k}[k_F,k_F]`$. The calculation of the energy is required to start from $`E=2t\mathrm{cos}k_p2tL{\displaystyle _{k_F}^{k_F}}\mathrm{cos}k\rho (k)𝑑k,`$ where the integration limit $`k_F`$ is required to fulfill $`_{k_F}^{k_F}\rho (k)𝑑k=(N1)/L`$. By introducing $`\rho (k)=\rho _0(k)+\rho _1(k)/L`$ etc., the excitation energy $`\mathrm{\Delta }E=EE_0`$ is composed of two terms $$\mathrm{\Delta }E(\overline{k},k_p)=\epsilon _h(\overline{k})+\overline{\epsilon }_h(k_p).$$ (134) Here we introduced the holon energy $`\epsilon _h(x)=2t\mathrm{cos}x{\displaystyle _{k_F}^{k_F}}(2t\mathrm{cos}k+\mathrm{\Lambda })\rho _1^c(k,x)𝑑k,`$ (135) and antiholon (“particle” state) energy $`\overline{\epsilon }_h(x)=\epsilon _h(x)`$. In eq.(135), the $`\rho _1^c(k,x)`$ should be solved from the following equation $`\rho _1^c(k,x)+{\displaystyle \frac{\mathrm{cos}k}{c}}R_{3/2}\left({\displaystyle \frac{\mathrm{sin}k\mathrm{sin}x}{c}}\right)=`$ (136) $`{\displaystyle \frac{\mathrm{cos}k}{c}}{\displaystyle _{k_F}^{k_F}}R_{3/2}\left({\displaystyle \frac{\mathrm{sin}ksink^{}}{c}}\right)\rho _1^c(k^{},x)𝑑k^{}.`$ (137) This equation is derived from (116) by taking $`\rho _1^{(o)}(k)`$ $`=`$ $`\delta (k\overline{k}),`$ (138) $`\sigma _1^{(o)}(\lambda )`$ $`=`$ $`K_{1/2}(\lambda \mathrm{sin}k_p),`$ (139) $`\omega _1^{(o)}(\mu )`$ $`=`$ $`\tau _1^{(o)}(\nu )=0,`$ (140) which comes from the “particle” and the hole. Now we find that this kind of excitation consists of a holon carrying energy $`\epsilon (\overline{k})`$ and an antiholon carrying energy $`\overline{\epsilon }_h(k_p)`$. Obviously, eq.(134) vanishes when $`\overline{k}k_F`$ and $`k_pk_F`$, and the holon-antiholon excitation is gapless. ### B The holon-holon excitation In order to discuss states with double occupancy, we need to consider solutions containing complex $`k`$ pairs. Suppose a configuration $`\{h_j\}`$ results in a complex pairs (two-strings), $`k^\pm =\kappa i\chi `$ and two holes inside the Fermi-sea, $`\overline{k}_1`$, $`\overline{k}_2[k_F,k_F]`$. After a careful analyses of the Bethe-ansatz equation, one finds the string position is restricted to lie around a particular $`\lambda `$ solution that we denoted by $`\lambda _0`$, i.e., it must satisfy $`\mathrm{sin}(\kappa \pm i\chi )=\lambda _0ic/2+O(e^{\eta L}).`$ An exact deletion holds in the second equation of (22) for $`\lambda _0=(\mathrm{sin}\overline{k}_1+\mathrm{sin}\overline{k}_2)/2`$. After rewriting the Bethe-ansatz equation by separating the factors of the complex $`k`$ pairs, we take the thermodynamics limit as before. In order to get the excitation energy we need to solve (116) with $`\rho _1^{(o)}(k)`$ $`=`$ $`\delta (k\overline{k}_1)+\delta (k\overline{k}_2)`$ (141) $`+\mathrm{cos}kK_{1/2}(\mathrm{sin}k\lambda _0),`$ (142) $`\sigma _1^{(o)}(\lambda )`$ $`=`$ $`\tau _1^{(o)}(\nu )=0,`$ (143) $`\omega _1^{(o)}(\mu )`$ $`=`$ $`K_{1/2}(\mu \lambda _0).`$ (144) After careful calculation, we obtain the excitation energy $`\mathrm{\Delta }E(\overline{k}_1,\overline{k}_2)=\epsilon _h(\overline{k}_1)+\epsilon _h(\overline{k}_1)+\mathrm{\Delta }(U,k_F),`$ where $`\epsilon _h`$ is the holon energy given by the same equation (135), and $`\mathrm{\Delta }(U,k_F)`$ is given by $`\mathrm{\Delta }(U,k_F)=U+2t{\displaystyle _\pi ^\pi }\mathrm{cos}^2kK_1(\mathrm{sin}k\lambda _0)𝑑k`$ (145) $`{\displaystyle _{k_F}^{k_F}}(2t\mathrm{cos}k+\mathrm{\Lambda })\rho _1^{stri}(k,\lambda _0)𝑑k,`$ (146) with $`\rho _1^{stri}(k,\lambda _0)={\displaystyle \frac{\mathrm{cos}k}{c}}R_1\left({\displaystyle \frac{\mathrm{sin}k\lambda _0}{c}}\right)`$ (147) $`+`$ $`{\displaystyle \frac{\mathrm{cos}k}{c}}{\displaystyle _{k_F}^{k_F}}R_{3/2}\left({\displaystyle \frac{\mathrm{sin}ksink^{}}{c}}\right)\rho _1^{stri}(k^{},\lambda _0)𝑑k^{}`$ (148) $`+`$ $`{\displaystyle \frac{\mathrm{cos}k}{c}}K_{1/2}(\mathrm{sin}k\lambda _0).`$ (149) Clearly the holon-holon excitation always has a gap $`\mathrm{\Delta }_g=2\epsilon _h(k_F)+\mathrm{\Delta }(U,k_F)`$, which exists at any filling. However, the gapless modes of holon-antiholon are available to carry charges for away from quarter-filling. It is easy to shown by calculating (103) that both holon-antiholon and holon-holon excitations are SU(4) singlet, consequently, they carry neither spins nor orbitals. ## IX Discussions In the above we have presented an extensive discussion on one-dimensional Hubbard-like model with SU(4) symmetry, where the sites are restricted to be occupied by at most two electrons.. The model was proposed to describe electrons with two-fold orbital degeneracy. The symmetries and some general features were given previously . We focused on the one dimensional case in this paper and studied the ground state and excitations by means of an exact solution. The excitation energies of the excited states are just sums of some particular terms related to quasiparticles. It provides an explicitly interpretation of the separation of charge excitations and spin-orbital excitations. Among the charge excitations, there are gapless holon-antiholon excitations and holon-holon excitations with gap. Both excitations carry neither spins nor orbitals. They are completely decoupled from the spin-orbital degree of freedom. The holons and antiholons move throughout the crystal at less than quarter-filling. Various excitations in spin-orbital sector consist of three basic modes which are created by the holes in the three rapidities for the spin-orbital double. That means there are three kinds of quasiparticles that carry spins and orbitals, i.e., two quadruplets transforming according to the fundamental or conjugate representation, and one hexaplet forming the six-dimensional representation. These elementary excitations in spin-orbital sector are gapless. As the on-site coupling in our model is isotropic for spin-orbital labels, there is no separation between spin and orbital. A complete separation between the spin waves and orbital waves is expected to occur after taking account of the contributions of Hund’s rule. This needs to introduce anisotropic on-site coupling in the spin and orbital configuration. For finite $`N`$ and $`L`$ we plotted the excitation spectra by solving the Bethe-ansatz equation numerically. The variation of the quantum number for excited states from that for the ground state, and their changes from integer to half-integers (or vice versa) were shown in each cases. It provides a concrete interpretation about the collective excitations for the orbital degenerate electronic systems. The overall structure of the spectra for spin-orbital excitations changed greatly with respect to the changes of the correlation strength. The lowest excitation energy and the whole pattern are raised when the correlation strength decreased. However, the “particle”-hole excitation spectrum does not change much from the strong to the weak correlation strengths. In the quarter-filled band for strong repulsive on-site coupling, there will be no doubly occupied sites. In this case the total wave function will be separated into a product of Slater determinant of $`N`$ “spinless” fermions and part of SU(4) Heisenberg magnets. The direct results from the Bethe-ansatz equation by taking strong coupling limit agree with it exactly. It is worthwhile to mention that the model studied here is not a direct SU(4) generalization of the Hubbard model, since a projection onto the subspace of states having at most two electrons at each site was made to render it solvable through the Bethe-ansatz. The two models are therefore not expected to share the same physical features. Considering the self-conjugate representation on a bipartite lattice in the strong repulsive coupling limit, Ref. clarified the system is dimerized with doubly degenerate singlet ground state and indicated the excitations are massive symmetry and antisymmetric kinks. In our present model, however, the local states on each site carry out the fundamental representation of SU(4). YQL acknowledges the supports of AvH Stiftung and interesting discussions with H Frahm. This work is supported by NSFC-19975040 and EYF of China Education Ministry.
warning/0001/hep-th0001129.html
ar5iv
text
# Algebraic Quantum Field Theory, Perturbation Theory, and the Loop Expansion ## 1 Introduction Quantum field theory is a very successful frame for our present understanding of elementary particle physics. In the case of QED it led to fantastically precise predictions of experimentally measurable quantities; moreover the present standard model of elementary particle physics is of a similar structure and is also in good agreement with experiments. Unfortunately, it is not so clear what an interacting quantum field theory really is, expressed in meaningful mathematical terms. In particular, it is by no means evident how the local algebras of observables can be defined. A direct approach by methods of constructive field theory led to the paradoxical conjecture that QED does not exist; the situation seems to be better for Yang-Mills theories because of asymptotic freedom, but there the problem of big fields which can appear at large volumes poses at present unsurmountable problems . In this paper we will take a pragmatic point of view: interacting quantum field theory certainly exists on the level of perturbation theory, and our confidence on quantum field theory relies mainly on the agreement of experimental data with results from low orders of perturbation theory. On the other hand, the general structure of algebraic quantum field theory (or ’local quantum physics’) coincides nicely with the qualitative features of elementary particle physics, therefore it seems to be worthwhile to revisit perturbation theory from the point of view of algebraic quantum field theory. This will, on the one hand side, provide physically relevant examples for algebraic quantum field theory, and on the other hand, give new insight into the structure of perturbation theory. In particular, we will see, that we can reach a complete separation of the infrared problem from the ultraviolet problem. This might be of relevance for Yang-Mills theory, and it is important for the construction of the theory on curved spacetimes . The plan of the paper is as follows. We will start by describing the Stückelberg-Bogoliubov-Shirkov-Epstein-Glaser-version of perturbation theory . This construction yields the local $`S`$-matrices $`S(g)(g𝒟(^4))`$ as formal power series in $`g`$ (Sect. 2). The most important requirement which is used in this construction is the condition of causality (15) which is a functional equation for $`gS(g)`$. The results of Sects. 3 and 4 are to a large extent valid beyond perturbation theory. We only assume that we are given a family of unitary solutions of the condition of causality. In terms of these local $`S`$-matrices we will construct nets of local observable algebras for the interacting theory (sect. 3). We will see that, as a consequence of causality, the interacting theory is completely determined if it is known for arbitrary small spacetime volumes (Sect. 4). In Sect. 5 we algebraically quantize a free field by deforming the (classical) Poisson algebra. In a second step we generalize this quantization procedure to the perturbative interacting field. We end up with an algebraic formulation of the expansion in $`\mathrm{}`$ of the interacting observables (’loop expansion’). In the last section we investigate two examples for the quantum action principle: the field equation and the variation of a parameter in the interaction. Usually this principle is formulated in terms of Green’s functions , i.e. the vacuum expectation values of timeordered products of interacting fields. Here we give a local algebraic formulation, i.e. an operator identity for a localized interaction. In the case of the variation of a parameter in the interaction this requires the use of the retarded product of interacting fields, instead of only time ordered products (as in the formulation in terms of Green’s functions). For a local construction of observables and physical states in gauge theories we refer to . There, perturbative positivity (“unitarity”) is, by a local version of the Kugo-Ojima formalism , reduced to the validity of BRST symmetry . ## 2 Free fields, Borchers’ class and local $`S`$-matrices An algebra of observables corresponding to the Klein-Gordon equation $$(\text{ }\text{ }\text{ }\text{ }+m^2)\phi =0$$ (1) can be defined as follows: Let $`\mathrm{\Delta }_{\mathrm{ret},\mathrm{av}}`$ be the retarded, resp. advanced Green’s functions of $`(\text{ }\text{ }\text{ }\text{ }+m^2)`$ $$(\text{ }\text{ }\text{ }\text{ }+m^2)\mathrm{\Delta }_{\mathrm{ret},\mathrm{av}}=\delta ,\mathrm{supp}\mathrm{\Delta }_{\mathrm{ret},\mathrm{av}}\overline{V}_\pm ,$$ (2) where $`\overline{V}_\pm `$ denotes the closed forward, resp. backward lightcone, and let $`\mathrm{\Delta }=\mathrm{\Delta }_{\mathrm{ret}}\mathrm{\Delta }_{\mathrm{av}}`$. The algebra of observables $`𝒜`$ is generated by smeared fields $`\phi (f),f𝒟(^4)`$, which obey the following relations $`f\phi (f)\mathrm{is}\mathrm{linear},`$ (3) $`\phi ((\text{ }\text{ }\text{ }+m^2)f)=0,`$ (4) $`\phi (f)^{}=\phi (\overline{f}),`$ (5) $`[\phi (f),\phi (g)]=i<f,\mathrm{\Delta }g>.`$ (6) where the star denotes convolution and $`<f,g>=d^4xf(x)g(x)`$. As a matter of fact, $`𝒜`$ (as a $``$-algebra with unit) is uniquely determined by these relations. The Fock space representation $`\pi `$ of the free field is induced via the GNS-construction from the vacuum state $`\omega _0`$. Namely, let $`\omega _0:𝒜`$ be the quasifree state given by the two-point function $$\omega _0(\phi (f)\phi (g))=i<f,\mathrm{\Delta }_+g>$$ (7) where $`\mathrm{\Delta }_+`$ is the positive frequency part of $`\mathrm{\Delta }`$. Then the Fock space $``$, the vector $`\mathrm{\Omega }`$ representing the vacuum and the Fock representation are up to equivalence determined by the relation $$(\mathrm{\Omega },\pi (A)\mathrm{\Omega })=\omega _0(A),A𝒜.$$ On $``$, the field $`\phi `$ (we will omit the representation symbol $`\pi `$) is an operator valued distribution, i.e. there is some dense subspace $`𝒟`$ with $`(i)\phi (f)\mathrm{End}(𝒟)`$ $`(ii)f\phi (f)\mathrm{\Phi }\mathrm{is}\mathrm{continuous}\mathrm{\Phi }𝒟.`$ There are other fields $`A`$ on $``$, on the same domain, which are relatively local to $`\phi `$, $$[A(f),\phi (g)]=0\mathrm{if}(xy)^2<0(x,y)(\mathrm{supp}f\times \mathrm{supp}g).$$ (8) They form the so called Borchers class $``$. In the case of the free field in 4 dimensions, $``$ consists of Wick polynomials and their derivatives . Fields from the Borchers class can be used to define local interactions, $$H_I(t)=d^3xg(t,\stackrel{}{x})A(t,\stackrel{}{x}),g𝒟(^4),$$ (9) (where the minus sign comes from the interpretation of $`A`$ as an interaction term in the Lagrangian) provided they can be restricted to spacelike surfaces. The corresponding time evolution operator from $`\tau `$ to $`\tau `$, where $`\tau >0`$ is so large that $`\mathrm{supp}g(\tau ,\tau )\times \mathrm{R}^3`$, (the $`S`$-matrix) is formally given by the Dyson series $$S(g)=\mathrm{𝟏}+\underset{n=1}{\overset{\mathrm{}}{}}\frac{i^n}{n!}𝑑x_1\mathrm{}𝑑x_nT\left(A(x_1)\mathrm{}A(x_n)\right)g(x_1)\mathrm{}g(x_n).$$ (10) with the time ordered products (’$`T`$-products’) $`T\left(\mathrm{}\right)`$. It is difficult to derive (10) from (9) if the field $`A`$ cannot be restricted to spacelike surfaces. Unfortunately, this is almost always the case in four spacetime dimensions, the only exception being the field $`\phi `$ itself and its derivatives. Therefore one defines the timeordered products of $`n`$ factors directly as multilinear (with respect to $`C^{\mathrm{}}`$-functions as coefficients) symmetric mappings from $`^n`$ to operator valued distributions $`T\left(A_1(x_1)\mathrm{}A_n(x_n)\right)`$ on $`𝒟`$ such that they satisfy the factorization condition<sup>1</sup><sup>1</sup>1Due to the symmetry and linearity of $`T(\mathrm{})`$ it suffices to consider the case $`A_1=A_2=\mathrm{}=A_n`$. $$T\left(A(x_1)\mathrm{}A(x_n)\right)=T\left(A(x_1)\mathrm{}A(x_k)\right)T\left(A(x_{k+1})\mathrm{}A(x_n)\right)$$ (11) if $`\{x_{k+1},\mathrm{},x_n\}(\{x_1,\mathrm{},x_k\}+\overline{V}_+)=\mathrm{}`$. The $`S`$-matrix $`S(g)`$ is then, as a formal power series, by definition given by (10) . Since its zeroth order term is $`\mathrm{𝟏}`$, it has an inverse in the sense of formal power series $$S(g)^1=\mathrm{𝟏}+\underset{n=1}{\overset{\mathrm{}}{}}\frac{(i)^n}{n!}𝑑x_1\mathrm{}𝑑x_n\overline{T}\left(A(x_1)\mathrm{}A(x_n)\right)g(x_1)\mathrm{}g(x_n),$$ (12) where the ’antichronological products’ $`\overline{T}(\mathrm{})`$ can be expressed in terms of the time ordered products $$\overline{T}\left(A(x_1)\mathrm{}A(x_n)\right)\stackrel{\mathrm{def}}{=}\underset{P𝒫(\{1,\mathrm{},n\})}{}(1)^{|P|+n}\underset{pP}{}T(A(x_i),ip).$$ (13) Here ($`𝒫(\{1,\mathrm{},n\})`$ is the set of all ordered partitions of $`\{1,\mathrm{},n\}`$ and $`|P|`$ is the number of subsets in $`P`$). The $`\overline{T}`$-products satisfy anticausal factorization $$\overline{T}\left(A(x_1)\mathrm{}A(x_n)\right)=\overline{T}\left(A(x_{k+1})\mathrm{}A(x_n)\right)\overline{T}\left(A(x_1)\mathrm{}A(x_k)\right)$$ (14) if $`\{x_{k+1},\mathrm{},x_n\}(\{x_1,\mathrm{},x_k\}+\overline{V}_+)=\mathrm{}`$. The crucial observation now (cf. ) is that $`S(g)`$ satisfies the remarkable functional equation $$S(f+g+h)=S(f+g)S(g)^1S(g+h),$$ (15) $`f,g,h𝒟(^4)`$, whenever $`(\mathrm{supp}f+\overline{V}_+)\mathrm{supp}h=\mathrm{}`$ (independent of $`g`$). Equivalent forms of this equation play an important role in and . For $`g=0`$ this is just the functional equation for the time evolution and may be interpreted as the requirement of causality . Actually, for formal power series $`S()`$ of operator valued distributions, the $`g=0`$ equation is equivalent to the seemingly stronger relation (15), because both are equivalent to condition (11) for the time ordered products. We call (15) the ’condition of causality’. ## 3 Interacting local nets The arguments of this and the next section are to a large extent independent of perturbation theory. We start from the assumption that we are given a family of unitaries $`S(f)𝒜`$, $`f𝒟(^4,𝒱)`$ (i.e. $`f`$ has the form $`f=_if_i(x)A_i,f_i𝒟(^4,),A_i𝒱`$) where $`𝒱`$ is an abstract, finite dimensional, real vector space, interpreted as the space of possible interaction Lagrangians, and $`𝒜`$ is some unital $``$-algebra. In perturbation theory $`𝒱`$ is a real subspace of the Borchers’ class. The unitaries $`S(f)`$ are required to satisfy the causality condition (15). We first observe that we obtain new solutions of (15) by introducing the relative $`S`$-matrices $$S_g(f)\stackrel{\mathrm{def}}{=}S(g)^1S(g+f),$$ (16) where now $`g`$ is kept fixed and $`S_g(f)`$ is considered as a functional of $`f`$. In particular, the relative $`S`$-matrices satisfy local commutation relations $$[S_g(h),S_g(f)]=0\mathrm{if}(xy)^2<0(x,y)\mathrm{supp}h\times \mathrm{supp}f.$$ (17) Therefore their functional derivatives $`A_g(x)=\frac{\delta }{\delta h(x)}S_g(hA)|_{h=0}`$, $`A𝒱`$, $`h𝒟(^4)`$, provided they exist, are local fields (in the limit $`g`$ constant this is Bogoliubov’s definition of interactig fields) . We now introduce local algebras of observables by assigning to a region $`𝒪`$ of Minkowski space the $``$-algebra $`𝒜_g(𝒪)`$ which is generated by $`\{S_g(h),h𝒟(𝒪,𝒱)\}`$. A remarkable consequence of relation (15) is that the structure of the algebra $`𝒜_g(𝒪)`$ depends only locally on $`g`$ , namely, if $`gg^{}`$ in a neighbourhood of a causally closed region containing $`𝒪`$, then there exists a unitary $`V𝒜`$ such that $$VS_g(h)V^1=S_g^{}(h),h𝒟(𝒪,𝒱).$$ (18) Hence the system of local algebras of observables (according to the principles of algebraic quantum field theory this system (“the local net”) contains the full physical content of a quantum field theory) is completely determined if one knows the relative $`S`$-matrices for test functions $`g𝒟(^4,𝒱)`$. The construction of the global algebra of observables for an interaction Lagrangian $`𝒱`$ may be performed explicitly (cf. ). Let $`\mathrm{\Theta }(𝒪)`$ be the set of all functions $`\theta 𝒟(^4)`$ which are identically to $`1`$ in a causally closed open neighbourhood of $`𝒪`$ and consider the bundle $$\underset{\theta \mathrm{\Theta }(𝒪)}{}\{\theta \}\times 𝒜_\theta (𝒪).$$ (19) Let $`𝒰(\theta ,\theta ^{})`$ be the set of all unitaries $`V𝒜`$ with $$VS_\theta (h)=S_\theta ^{}(h)V,h𝒟(𝒪,𝒱).$$ (20) Then $`𝒜_{}(𝒪)`$ is defined as the algebra of covariantly constant sections, i.e. $`𝒜_{}(𝒪)A=(A_\theta )_{\theta \mathrm{\Theta }(𝒪)}(A_\theta 𝒜_\theta (𝒪))`$ (21) $`VA_\theta =A_\theta ^{}V,V𝒰(\theta ,\theta ^{}).`$ (22) $`𝒜_{}(𝒪)`$ contains in particular the elements $`S_{}(h)`$, $$(S_{}(h))_\theta =S_\theta (h).$$ (23) The construction of the local net is completed by fixing the embeddings $`i_{21}:𝒜_{}(𝒪_1)𝒜_{}(𝒪_2)`$ for $`𝒪_1𝒪_2`$. But these embeddings are inherited from the inclusions $`𝒜_\theta (𝒪_1)𝒜_\theta (𝒪_2)`$ for $`\theta \mathrm{\Theta }(𝒪_2)`$ by restricting the sections from $`\mathrm{\Theta }(𝒪_1)`$ to $`\mathrm{\Theta }(𝒪_2)`$. The embeddings evidently satisfy the compatibility relation $`i_{12}i_{23}=i_{13}`$ for $`𝒪_3𝒪_2𝒪_1`$ and define thus an inductive system. Therefore, the global algebra can be defined as the inductive limit of local algebras $$𝒜_{}\stackrel{\mathrm{def}}{=}_𝒪𝒜_{}(𝒪).$$ (24) In perturbation theory, the unitaries $`V𝒰(\theta ,\theta ^{})`$ are themselves formal power series, therefore it makes no sense to say that two elements $`A,B𝒜_{}(𝒪)`$ agree in $`n`$-th order, but only that they agree up to $`n`$-th order (because $`(A_\theta B_\theta )=𝒪(g^{n+1})`$ implies $`A_\theta ^{}B_\theta ^{}=V^1(A_\theta B_\theta )V=𝒪(g^{n+1})`$). The time ordered products and hence the relative $`S`$-matrices $`S_\theta (h)`$ are chosen as to satisfy Poincaré covariance (see the normalization condition N1 below), i.e. the unitary positive energy representation $`U`$ of the Poincaré group $`𝒫_+^{}`$ under which the free field transforms satisfies $$U(L)S_\theta (h)U(L)^1=S_{\theta _L}(h_L),\theta _L(x):=\theta (L^1x),h_L(x):=D(L)h(L^1x),$$ (25) $`L𝒫_+^{}`$ provided $``$ is a Lorentz scalar and $`𝒱`$ transforms under the finite dimensional representation $`D`$ of the Lorentz group. This enables us to define an automorphic action of the Poincaré group on the algebra of observables. Let for $`A𝒜_{}(𝒪),\theta \mathrm{\Theta }(L𝒪)`$ $$(\alpha _L(A))_\theta \stackrel{\mathrm{def}}{=}U(L)A_{\theta _{L^1}}U(L)^1.$$ (26) By inserting the definitions one finds that $`\alpha _L(A)`$ is again a covariantly constant section (22). So $`\alpha _L`$ is an automorphism of the net which realizes the Poincaré symmetry $$\alpha _L𝒜_{}(𝒪)=𝒜_{}(L𝒪),\alpha _{L_1L_2}=\alpha _{L_1}\alpha _{L_2}.$$ (27) For the purposes of perturbation theory, we have to enlarge the local algebras somewhat. In perturbation theory, the relative S-matrices are formal power series in two variables, and therefore the generators of the local algebras $$S_{}(\lambda f)=\underset{n=0}{\overset{\mathrm{}}{}}\frac{i^n\lambda ^n}{n!}T_{}(f^n)$$ (28) are formal power series with coefficients which are covariantly constant sections in the sense of (22). The first order terms in (28) are, according to Bogoliubov, the interacting local fields, $$T_{}(hA)=:A_{}(h),A𝒱,h𝒟(^4),$$ (29) the higher order terms satisfy the causality condition (11) and may therefore be interpreted as time ordered products of interacting fields (cf. sect. 8.1) Our enlarged local algebra $`𝒜_{}(𝒪)`$ (we use the same symbol as before) now consists of all formal power series with coefficients from the algebra generated by all timeordered products $`T_{}(f^n)`$ with $`f𝒟(𝒪,𝒱),n_0`$. ## 4 Consequences of causality Another consequence of the causality relation (15) is that the $`S`$-matrices $`S(f)`$ are uniquely fixed if they are known for test functions with arbitrarily small supports. Namely, by a repeated use of (15) we find that $`S(_{i=1}^nf_i)`$ is a product of factors $`S(_{iK}f_i)^{\pm 1}`$ where the sets $`K\{1,\mathrm{},n\}`$ have the property that for every pair $`i,jK`$ the causal closures of $`\mathrm{supp}f_i`$ and $`\mathrm{supp}f_j`$ overlap. Hence if the supports of all $`f_i`$ are contained in double cones of diameter $`d`$, the supports of $`_{iK}f_i`$ fit into double cones of diameter $`2d`$. As $`d>0`$ can be chosen arbitrarily small and the relative $`S`$-matrices also satisfy (15), this implies additivity of the net, $$𝒜_{}(𝒪)=\underset{\alpha }{}𝒜_{}(𝒪_\alpha )$$ (30) where $`(𝒪_\alpha )`$ is an arbitrary covering of $`𝒪`$ and where the symbol $``$ means the generated algebra. One might also pose the existence question: Suppose we have a family of unitaries $`S(f)`$ for all $`f`$ with sufficiently small support which satisfy the causality condition (15) for $`f,g,h𝒟(𝒪,𝒱)`$, diam$`(𝒪)`$ sufficiently small, and local commutativity for arbitrary big separation $$[S(f),S(g)]=0\text{if}\mathrm{supp}f\text{is spacelike to }\mathrm{supp}g.$$ By repeated use of the causality (15) we can then define $`S`$-matrices for test functions with larger support. It is, however, not evident that these $`S`$-matrices are independent of the way of construction and that they satisfy the causality condition. (We found a consistent construction only in the simple case of one dimension: $`x=`$ time.) Fortunately, a general positive answer can be given in perturbation theory. Let $`S(f)`$ be given for $`f𝒟(𝒪,𝒱)`$ for all double cones with diam$`(𝒪)<r`$. The time ordered product of $`n`$ factors is the $`n`$-fold functional derivative of $`S`$ at $`f=0`$. It is an operator valued distribution<sup>2</sup><sup>2</sup>2Here we change the notation for the time ordered products: let $`f=_if_i(x)A_i,f_i𝒟(^4),A_i𝒱`$. Instead of $`𝑑x_1\mathrm{}𝑑x_n_{i_1\mathrm{}i_n}T\left(A_{i_1}(x_1)\mathrm{}A_{i_n}(x_n)\right)f_{i_1}(x_1)\mathrm{}f_{i_n}(x_n)`$ (10) we write $`𝑑x_1\mathrm{}𝑑x_nT_n(x_1,\mathrm{},x_n)f(x_1)\mathrm{}f(x_n)T_n(f^n)`$. $`T_n`$ defined on test functions of $`n`$ variables with support contained in $`𝒰_n\stackrel{\mathrm{def}}{=}\{(y_1,\mathrm{},y_n)^{4n}|\mathrm{max}_{i<j}|y_iy_j|<\frac{r}{2}\}`$ and with values in $`𝒱^n`$. Especially we know $`T_1(x)`$ on $`^4`$. On this domain the time ordered products satisfy the factorization condition (11). In addition, local commutativity of the $`S`$-matrices implies $$[T_n(x_1,\mathrm{},x_n),T_m(y_1,\mathrm{}y_m)]=0$$ (31) for $`(x_iy_j)^2<0(i,j)`$ and $`(x_1,\mathrm{}x_n)𝒰_n,(y_1,\mathrm{},y_m)𝒰_m`$. By construction $`T_n|_{𝒰_n}`$ is symmetric with respect to permutations of the factors. We now show that this input suffices to construct $`T_n(x_1,\mathrm{},x_n)`$ on the whole $`^{4n}`$ by induction on $`n`$. We assume that the $`T_k`$’s were constructed for $`kn1`$, that they fulfil causality (11) and $$[T_m(x_1,\mathrm{},x_m),T_k(y_1,\mathrm{}y_k)]=0\mathrm{for}(x_1,\mathrm{}x_m)𝒰_m,kn1$$ (32) ($`m`$ arbitrary) and $$[T_l(x_1,\mathrm{},x_l),T_k(y_1,\mathrm{}y_k)]=0\mathrm{for}l,kn1,$$ (33) if $`(x_iy_j)^2<0(i,j)`$ in the latter two equations. We can now proceed as in Sect. 4 of . <sup>3</sup><sup>3</sup>3In contrast to the (inductive) Epstein-Glaser construction of $`T_n(x_1,\mathrm{},x_n)`$ the present construction is unique, normalization conditions (e.g. $`\mathrm{𝐍𝟏}\mathrm{𝐍𝟒}`$ in sect. 5) are not needed, because the non-uniqueness of the Epstein-Glaser construction is located at the total diagonal $`\mathrm{\Delta }_n\{(x_1,\mathrm{},x_n)|x_1=\mathrm{}=x_n\}`$. But here the time ordered products are given in the neighbourhood $`𝒰_n`$ of $`\mathrm{\Delta }_n`$. Let $`𝒥`$ denote the family of all non-empty proper subsets $`I`$ of the index set $`\{1,\mathrm{},n\}`$ and define the sets $`𝒞_I\stackrel{\mathrm{def}}{=}\{(x_1,\mathrm{},x_n)^{4n}|x_iJ^{}(x_j),iI,jI^c\}`$ for any $`I𝒥`$. Then $$\underset{I𝒥}{}𝒞_I𝒰_n=^{4n}.$$ (34) We use the short hand notations $$T^I(x_I)=T(\underset{iI}{}A_i(x_i)),x_I=(x_i,iI).$$ (35) On $`𝒟(𝒞_I)`$ we set $$T_I(x)\stackrel{\mathrm{def}}{=}T^I(x_I)T^{I^c}(x_{I^c})$$ (36) for any $`I𝒞_I`$. For $`I_1,I_2𝒥,𝒞_{I_1}C_{I_2}\mathrm{}`$ one easily verifies<sup>4</sup><sup>4</sup>4In contrast to the Wick expansion of the $`T`$-products is not used here, because local commutativity of the $`T`$-products is contained in the inductive assumption. $$T_{I_1}|_{𝒞_{I_1}𝒞_{I_2}}=T_{I_2}|_{𝒞_{I_1}𝒞_{I_2}}.$$ (37) Let now $`\{f_I\}_{I𝒥}\{f_0\}`$ be a finite smooth partition of unity of $`^{4n}`$ subordinate to $`\{𝒞_I\}_{I𝒥}𝒰_n`$: $`\mathrm{supp}f_I𝒞_I,\mathrm{supp}f_0𝒰_n`$. Then we define $$T_n(h)\stackrel{\mathrm{def}}{=}T_n|_{𝒰_n}(f_0h)+\underset{I𝒥}{}T_I(f_Ih),h𝒟(^{4n},𝒱^n).$$ (38) As in one may prove that this definition is independent of the choice of $`\{f_I\}_{I𝒥}\{f_0\}`$ and that $`T_n`$ is symmetric with respect to permutations of the factors and satisfies causality (11). Local commutativity (32) and (33) (with $`n1`$ replaced by $`n`$) is verified by inserting the definition (38) and using the assumptions. By (10) we obtain from the $`T`$-products the corresponding $`S`$-matrix $`S(g)`$ for arbitrary large support of $`g𝒟(^4,𝒱)`$, and $`S(g)`$ satisfies the functional equation (15). ## 5 Perturbative quantization and loop expansion Causal perturbation theory was traditionally formulated in terms of operator valued distributions on Fock space. It is therefore well suited for describing the deformation of the free field into an interacting field by turning on the interaction $`g𝒟(^4,𝒱)`$. It is much less clear how an expansion in powers of $`\mathrm{}`$ can be performed, describing the deformation of the classical field theory, mainly because the Fock space has no classical counter part. Usually the expansion in powers of $`\mathrm{}`$ is done in functional approaches to field theory by ordering Feynman graphs according to loop number. In this section we show that the algebraic description provides a natural formulation of the loop expansion, and we point out the connection to formal quantization theory. ### 5.1 Quantization of a free field and Wick products In quantization theory one associates to a given classical theory a quantum theory. One procedure is the deformation (or star-product) quantization . This procedure starts from a Poisson algebra, i.e. a commutative and associative algebra together with a second product: a Poisson bracket, satisfying the Leibniz rule and the Jacobi identity; and to deform the product as a function of $`\mathrm{}`$, such that<sup>5</sup><sup>5</sup>5The deformed product is called a $``$-product in deformation theory. In order to avoid confusion with the $``$-operation we denote the product by $`\times _{\mathrm{}}`$. $`a\times _{\mathrm{}}b`$ is a formal power series in $`\mathrm{}`$, the associativity is maintained and $$a\times _{\mathrm{}}b\stackrel{\mathrm{}0}{}ab,\frac{1}{\mathrm{}}(a\times _{\mathrm{}}bb\times _{\mathrm{}}a)\stackrel{\mathrm{}0}{}\{a,b\}.$$ (39) Actually this scheme can easily be realized in free field theory (cf. ). Basic functions are the evaluation functionals $`\phi _{\mathrm{class}}(x)`$, $`(\text{ }\text{ }\text{ }\text{ }+m^2)\phi _{\mathrm{class}}=0`$, with the Poisson bracket $$\{\phi _{\mathrm{class}}(x),\phi _{\mathrm{class}}(y)\}=\mathrm{\Delta }(xy)$$ (40) ($`\mathrm{\Delta }`$ is the commutator function (2)). Because of the singular character of $`\mathrm{\Delta }`$ the fields should be smoothed out in order to belong to the Poisson algebra. Hence our fundamental classical observables are $$\varphi (t)=t_0+\underset{n=1}{\overset{N}{}}\phi _{\mathrm{class}}(x_1)\mathrm{}\phi _{\mathrm{class}}(x_n)t_n(x_1,\mathrm{},x_n)𝑑x_1\mathrm{}𝑑x_n,t(t_0,t_1,\mathrm{}),$$ (41) where $`t_0`$ arbitrary, $`N<\mathrm{}`$, $`t_n`$ is a suitable test “function” (we will admit also certain distributions) with compact support. The Klein Gordon equation shows up in the property: $`A(t)=0`$ if $`t_0=0`$ and $`t_n=(\text{ }\text{ }\text{ }\text{ }_i+m^2)g_n`$ for all $`n>0`$, some $`i=i(n)`$ and some $`g_n`$ with compact support. In the quantization procedure we identify $`\phi _{\mathrm{class}}(x_1)\mathrm{}\phi _{\mathrm{class}}(x_n)`$ with the normally ordered product (Wick product) $`:\phi (x_1)\mathrm{}\phi (x_n):`$ ($`\phi `$ is the free quantum field (3-6)). Wick’s theorem may be interpreted as the definition of a $`\mathrm{}`$-dependent associative product, $`:{\displaystyle \underset{iI}{}}\phi (x_i):\times _{\mathrm{}}:{\displaystyle \underset{jJ}{}}\phi (x_j):`$ $`=`$ $`{\displaystyle \underset{KI}{}}{\displaystyle \underset{\alpha :KJ\mathrm{injective}}{}}{\displaystyle \underset{jK}{}}i\mathrm{}\mathrm{\Delta }_+(x_jx_{\alpha (j)})`$ $`:{\displaystyle \underset{l(IK)(J\alpha (K))}{}}\phi (x_l):`$ (42) in the linear space spanned by Wick products (the “Wick quantization”).<sup>6</sup><sup>6</sup>6The observation that the Wick quantization is appropriate for the quantization of the free field goes back to Dito . To be precise we have to fix a suitable test function space (or better: test distribution space) in (41) which is small enough such that the product is well defined for all $`\mathrm{}`$ and which contains the interesting cases occuring in perturbation theory, e.g. products of translation invariant distributions (particularly $`\delta `$-distributions of difference variables) with test functions of compact support should be allowed for $`t_n`$ as in Theorem 0 of Epstein and Glaser. Let $$𝒲_n\stackrel{\mathrm{def}}{=}\{t𝒟^{}(^{4n})_{\mathrm{symm}},\mathrm{supp}t\text{ compact },\mathrm{WF}(t)(^{4n}\times \overline{V_+^nV_{}^n})=\mathrm{}\}$$ (43) (see the Appendix for a definition of the wave front set $`\mathrm{WF}`$ of a distribution). In it was shown that Wick polynomials smeared with distributions $`t𝒲_n`$, $$(\phi ^n)(t)\stackrel{\mathrm{def}}{=}:\phi (x_1)\mathrm{}\phi (x_n):t(x_1,\mathrm{},x_n)dx_1\mathrm{}dx_n,(\phi ^0)\stackrel{\mathrm{def}}{=}\mathbf{\hspace{0.17em}1},$$ (44) are densely defined operators on an invariant domain in Fock space. This includes in particular the Wick powers $$:\phi ^n(f):=(\phi ^n)(t),f𝒟(^4),t(x_1,\mathrm{},x_n)=f(x_1)\underset{i=2}{\overset{n}{}}\delta (x_ix_1)$$ (45) The product of two such operators is given by $$(\phi ^n)(t)\times _{\mathrm{}}(\phi ^m)(s)=\underset{k=0}{\overset{\mathrm{min}\{n,m\}}{}}\mathrm{}^k(\phi ^{(n+m2k)})(t_ks)$$ (46) with the $`k`$-times contracted tensor product $`(t_ks)(x_1,\mathrm{},x_{n+m2k})=𝒮{\displaystyle \frac{n!m!i^k}{k!(nk)!(mk)!}}{\displaystyle 𝑑y_1\mathrm{}𝑑y_{2k}\mathrm{\Delta }_+(y_1y_2)\mathrm{}}`$ $`\mathrm{\Delta }_+(y_{2k1}y_{2k})t(x_1,\mathrm{},x_{nk},y_1,y_3,\mathrm{},y_{2k1})`$ $`s(x_{nk+1},\mathrm{},x_{n+m2k},y_2,y_4,\mathrm{},y_{2k})`$ (47) ($`𝒮`$ means the symmetrization in $`x_1,\mathrm{},x_{n+m2k}`$). The conditions on the wave front sets of $`t`$ and $`s`$ imply that the product $`(t_ks)`$ exists (see the Appendix) and is an element of $`𝒲_{n+m2k}`$. The $``$-operation reduces to complex conjugation of the smearing function. Let $`𝒲_0\stackrel{\mathrm{def}}{=}`$ and $`𝒲\stackrel{\mathrm{def}}{=}_{n=0}^{\mathrm{}}𝒲_n`$. For $`t𝒲`$ let $`t_n`$ denote the component of $`t`$ in $`𝒲_n`$. The $``$-operation is defined by $`(t^{})_n\stackrel{\mathrm{def}}{=}(\overline{t}_n)`$. Equation (46) can be thought of as the definition of an associative product on $`𝒲`$, $$(t\times _{\mathrm{}}s)_n=\underset{m+l2k=n}{}\mathrm{}^kt_m_ks_l.$$ (48) The Klein-Gordon equation defines an ideal $`𝒩`$ in $`𝒲`$ which is generated by $`(\text{ }\text{ }\text{ }\text{ }+m^2)f,f𝒟(^4)`$. Actually this ideal is independent of $`\mathrm{}`$ (because a contraction with $`(\text{ }\text{ }\text{ }\text{ }+m^2)f`$ vanishes) and coincides with the kernel of $`\varphi `$ defined in (41). Hence the product (48) is well defined on the quotient space $`\overline{𝒲}=𝒲/𝒩`$. For a given positive value of $`\mathrm{}`$, $`\overline{𝒲}`$ is isomorphic to the algebra generated by Wick products $`(\phi ^n)(t),t𝒲_n`$ (44). In the limit $`\mathrm{}0`$ we find $`\underset{\mathrm{}0}{lim}\varphi (t)\times _{\mathrm{}}\varphi (s)`$ $`=`$ $`\underset{\mathrm{}0}{lim}\varphi ({\displaystyle \underset{n}{}}\mathrm{}^nt_ns)`$ (49) $`=`$ $`\varphi (t_0s)=\varphi (t)\varphi (s)`$ (we set $`(t_ks)_n\stackrel{\mathrm{def}}{=}_{m+l=n}t_{m+k}_ks_{l+k}`$, cf. (47)), with the classical product $``$, and $$\underset{\mathrm{}0}{lim}\frac{1}{i\mathrm{}}[\varphi (t),\varphi (s)]_{\mathrm{}}=\varphi (t_1ss_1t)=\{\varphi (t),\varphi (s)\}$$ (50) with the classical Poisson bracket. Thus $`(\overline{𝒲},\times _{\mathrm{}})`$ provides a quantization of the given Poisson algebra of the classical free field $`\phi _{\mathrm{class}}`$ (40). We point out that we have formulated the algebraic structure of smeared Wick products without using the Fock space. The Fock representation is recovered, via the GNS construction, from the vacuum state $`\omega _0(t)=t_0`$. It is faithful for $`\mathrm{}0`$ but is one dimensional in the classical limit $`\mathrm{}=0`$. This illustrates the superiority of the algebraic point of view for a discussion of the classical limit. ### 5.2 Normalization conditions and retarded products To study the perturbative quantization of interacting fields we need some technical tools which are given in this subsection. The time ordered products are constructed by induction on the number $`n`$ of factors (which is also the order of the perturbation series (10)). In contrast to the inductive construction of the $`T`$-products in sect. 4, we do not know $`T_n|_{𝒰_n}`$ here. So causality (11) and symmetry determine the time ordered products uniquely (in terms of time ordered products of less factors) up to the total diagonal $`\mathrm{\Delta }_n=\{(x_1,\mathrm{},x_n)^{4n}|x_1=x_2=\mathrm{}=x_n\}`$. There is some freedom in the extension to $`\mathrm{\Delta }_n`$. To restrict it we introduce the following additional defining conditions (‘normalization conditions’, formulated for a scalar field without derivative coupling, i.e. $``$ is a Wick polynomial solely in $`\varphi `$, it does not contain derivatives of $`\varphi `$; for the generalization to derivative couplings see ) N1 covariance with resp. to Poincaré transformations and possibly discrete symmetries, in particular N2 unitarity: $`T(A_1(x_1)\mathrm{}A_n(x_n))^{}=\overline{T}(A_1^{}(x_1)\mathrm{}A_n^{}(x_n))`$, N3 $`[T(A_1(x_1)\mathrm{}A_n(x_n)),\varphi (x)]=`$ $`=i\mathrm{}_{k=1}^nT(A_1(x_1)\mathrm{}\frac{A_k}{\varphi }(x_k)\mathrm{}A_n(x_n))\mathrm{\Delta }(x_kx)`$, N4 $`(\text{ }\text{ }\text{ }\text{ }_x+m^2)T(A_1(x_1)\mathrm{}A_n(x_n)\varphi (x))=`$ $`=i\mathrm{}_{k=1}^nT(A_1(x_1)\mathrm{}\frac{A_k}{\varphi }(x_k)\mathrm{}A_n(x_n))\delta (x_kx)`$ where $`[\varphi (x),\varphi (y)]=i\mathrm{}\mathrm{\Delta }(xy)`$. N1 implies covariance of the arising theory, and N2 provides a $``$-structure. N3 gives the relation to time ordered products of sub Wick polynomials. Once these are known (in an inductive procedure), only a scalar distribution has to be fixed. Due to translation invariance the latter depends only on the relative coordinates. Hence, the extension of the (operator valued) $`T`$-product to $`\mathrm{\Delta }_n`$ is reduced to the extension of a C-number distribution $`t_0𝒟^{}(^{4(n1)}\{0\})`$ to $`t𝒟^{}(^{4(n1)})`$. (We call $`t`$ an extension of $`t_0`$ if $`t(f)=t_0(f),f𝒟(^{4(n1)}\{0\})`$). The singularity of $`t_0(y)`$ and $`t(y)`$ at $`y=0`$ is classified in terms of Steinmann’s scaling degree $$\mathrm{sd}(t)\stackrel{\mathrm{def}}{=}\mathrm{inf}\{\delta ,\underset{\lambda 0}{lim}\lambda ^\delta t(\lambda x)=0\}.$$ (51) By definition $`\mathrm{sd}(t_0)\mathrm{sd}(t)`$, and the possible extensions are restricted by requiring $$\mathrm{sd}(t_0)=\mathrm{sd}(t).$$ (52) Then the extension is unique for $`\mathrm{sd}(t_0)<4(n1)`$, and in the general case there remains the freedom to add derivatives of the $`\delta `$-distribution up to order $`(\mathrm{sd}(t_0)4(n1))`$, i.e. $$t(y)+\underset{|a|\mathrm{sd}(t_0)4(n1)}{}C_a^a\delta (y)$$ (53) is the general solution, where $`t`$ is a special extension , and the constants $`C_a`$ are restricted by N1, N2, N4, permutation symmetries and possibly further normalization conditions, e.g. the Ward identities for QED . For an interaction with mass dimension $`\mathrm{dim}()4`$ the requirement (52) implies renormalizability by power counting, i.e. the number of indeterminate constants $`C_a`$ does not increase by going over to higher perturbative orders. In it is shown that the normalization condition N4 implies the field equation for the interacting field corresponding to the free field $`\varphi `$ (see also (77) and sect. 6.1 below). We have defined the interacting fields as functional derivatives of relative S-matrices (29). Hence, to formulate the perturbation series of interacting fields we need the perturbative expansion of the relative S-matrices: $$S_g(f)=\underset{n,m}{}\frac{i^{n+m}}{n!m!}R_{n,m}(g^n;f^m),$$ (54) where $`g,f𝒟(^4,𝒱)`$. The coefficients are the so called retarded products (’$`R`$-products’). They can be expressed in terms of time ordered and anti-time ordered products by $$R_{n,m}(g^n;f^m)=\underset{k=0}{\overset{n}{}}(1)^k\frac{n!}{k!(nk)!}\overline{T}_k(g^k)\times _{\mathrm{}}T_{nk+m}(g^{(nk)}f^m).$$ (55) They vanish if one of the first $`n`$ arguments is not in the past light cone of some of the last $`m`$ arguments (, sect. 8.1), $$\mathrm{supp}R_{n,m}\left(\mathrm{}\right)\{(y_1,\mathrm{}y_n,x_1,\mathrm{},x_m),\{y_1,\mathrm{}y_n\}(\{x_1,\mathrm{},x_m\}+\overline{V}_{})\}.$$ (56) In the remaining part of this subsect. we show that the time ordered products can be defined in such a way that $`R_{n,m}`$ is of order $`\mathrm{}^n`$. For this purpose we will introduce the connected part $`(a_1\times _{\mathrm{}}\mathrm{}\times _{\mathrm{}}a_n)^c`$ of $`(a_1\times _{\mathrm{}}\mathrm{}\times _{\mathrm{}}a_n)`$, where the $`a_i`$ are normally ordered products of free fields, and the connected part $`T_n^c`$ of the time ordered product $`T_n`$ (or ’truncated time ordered product’). In both cases the connected part corresponds to the sum of connected diagrams, provided the vertices belonging to the same $`a_i`$ are identified. Besides the (deformed) product $`\times _{\mathrm{}}`$ (42) $$a\times _{\mathrm{}}b=\underset{n0}{}\mathrm{}^nM_n(a,b),$$ (57) where $`a,b`$ are normally ordered products of free fields, we have the classical product $`ab=M_0(a,b)`$, which is just the Wick product $$:\underset{iI}{}\phi (x_i)::\underset{jJ}{}\phi (x_j):=:\underset{iI}{}\phi (x_i)\underset{jJ}{}\phi (x_j):$$ (58) and which is also associative and in addition commutative. Then we define $`(a_1\times _{\mathrm{}}\mathrm{}\times _{\mathrm{}}a_n)^c`$ recursively by $$(a_1\times _{\mathrm{}}\mathrm{}\times _{\mathrm{}}a_n)^c\stackrel{\mathrm{def}}{=}(a_1\times _{\mathrm{}}\mathrm{}\times _{\mathrm{}}a_n)\underset{|P|2}{}\underset{JP}{}(a_{j_1}\times _{\mathrm{}}\mathrm{}\times _{\mathrm{}}a_{j_{|J|}})^c,$$ (59) where $`\{j_1,\mathrm{},j_{|J|}\}=J`$, $`j_1<\mathrm{}<j_{|J|}`$, the sum runs over all partitions $`P`$ of $`\{1,\mathrm{},n\}`$ in at least two subsets and $``$ means the classical product (58). $`T_n^c`$ is defined analogously $$T_n^c(f_1\mathrm{}f_n)\stackrel{\mathrm{def}}{=}T_n(f_1\mathrm{}f_n)\underset{|P|2}{}\underset{pP}{}T_{|p|}^c(_{jp}f_j),$$ (60) and similarly we introduce the connected antichronological product $`\overline{T}_n^c(\overline{T}_n)^c`$. Proposition 1: Let the normally ordered products of free fields $`a_1,\mathrm{},a_n`$ be of order $`𝒪(\mathrm{}^0)`$. Then $$(a_1\times _{\mathrm{}}\mathrm{}\times _{\mathrm{}}a_n)^c=𝒪(\mathrm{}^{n1}).$$ (61) Proof: We identify the vertices belonging to the same $`a_i`$ and apply Wick’s theorem (42) to $`a_1\times _{\mathrm{}}\mathrm{}\times _{\mathrm{}}a_n`$. Each ’contraction’ (i.e. each factor $`\mathrm{\Delta }_+`$) is accompanied by a factor $`\mathrm{}`$. In the terms $`\mathrm{}^0`$ (i.e. without any contraction) $`a_1,\mathrm{},a_n`$ are completely disconnected, the number of connected components is $`n`$. By a contraction this number is reduced by $`1`$ or $`0`$. So to obtain a connected term we need at least $`(n1)`$ contractions. Hence the connected terms are of order $`𝒪(\mathrm{}^{n1})`$. Let $`A_1,\mathrm{},A_n=𝒪(\mathrm{}^0)`$ and $`x_ix_j,1i<jn`$. Then there exists a permutation $`\pi 𝒮_n`$ such that $$T^c\left(A_1(x_1)\mathrm{}A_n(x_n)\right)=(A_{\pi 1}(x_{\pi 1})\times _{\mathrm{}}\mathrm{}\times _{\mathrm{}}A_{\pi n}(x_{\pi n}))^c=𝒪(\mathrm{}^{n1}).$$ (62) We want this estimate to hold true also for coinciding points $$T^c\left(A_1(x_1)\mathrm{}A_n(x_n)\right)=𝒪(\mathrm{}^{n1})\mathrm{on}𝒟(^{4n}).$$ (63) By the following argument this can indeed be satisfied by appropriate normalization of the time ordered products, i.e. (63) is an additional normalization condition, which is compatible with N1-N4. We proceed by induction on the number $`n`$ of factors. Let us assume that the $`T^c`$-products with less than $`n`$ factors fulfil (63) and that we are away from the total diagonal $`\mathrm{\Delta }_n`$. Using causal factorization, (60) and the shorthand notation $`T(J):=T(_{jJ}A_j(x_j)),J\{1,\mathrm{},n\}`$, we then know that there exists $`I\{1,\mathrm{},n\},I\mathrm{},I^c\mathrm{}`$, with $`T\left(A_1(x_1)\mathrm{}A_n(x_n)\right)=T(I)\times _{\mathrm{}}T(I^c)={\displaystyle \underset{r=1}{\overset{|I|}{}}}{\displaystyle \underset{s=1}{\overset{|I^c|}{}}}{\displaystyle \underset{I_1\mathrm{}I_r=I}{}}{\displaystyle \underset{J_1\mathrm{}J_s=I^c}{}}`$ $`{\displaystyle \underset{k0}{}}\mathrm{}^kM_k(T^c(I_1)\mathrm{}T^c(I_r),T^c(J_1)\mathrm{}T^c(J_s)),`$ (64) where $``$ means the disjoint union. We now pick out the connected diagrams. The term $`k=0`$ on the r.h.s. has $`(r+s)`$ disconnected components. Analogously to Proposition 1 we conclude that it must hold $`k(r+s1)`$ for a connected diagram. Taking the validity of (63) for $`T^c(I_l)`$ and $`T^c(J_m)`$ into account, we obtain $`_{l=1}^r(|I_l|1)+_{m=1}^r(|J_m|1)+(r+s1)=n1`$ for the minimal order in $`\mathrm{}`$ of a connected diagram. So the $`\mathrm{}`$-power behaviour (62) holds true on $`𝒟(^{4n}\mathrm{\Delta }_n)`$, and (63) is in fact a normalization condition. Due to (60) $`(T_nT_n^c)`$ is completely given by timeordered products of lower orders $`<n`$ and hence is known also on $`\mathrm{\Delta }_n`$. The problem of extending $`T_n`$ to $`\mathrm{\Delta }_n`$ concerns solely $`T_n^c`$. The normalization conditions N1 \- N4 are equivalent to the same conditions for $`T_n^c`$ and $`\overline{T}_n^c`$ (i.e. $`T_n`$ and $`\overline{T}_n`$ everywhere replaced by $`T_n^c`$ and $`\overline{T}_n^c`$). Due to N3 \- N4 it remains only the extension of $`<\mathrm{\Omega },T^c(A_1\mathrm{}A_n)\mathrm{\Omega }>`$ where all $`A_j`$ are different from free fields and $`\mathrm{\Omega }`$ is the vacuum. It is obvious that this can be done in a way which maintains (63) and is in accordance with N1 \- N2. We emphasize that the (ordinary) time ordered product $`T_n`$ does not satisfy (63) because of the presence of disconneted diagrams. On the other hand the connected antichronological product $`\overline{T}_n^c`$ fulfills the estimate (63), as may be seen by unitarity N2. We now turn to the retarded products (55): Proposition 2: Let $`𝒟(^4,𝒱)f_j,g_k=𝒪(\mathrm{}^0)`$. Then the following statements hold true: (i) All diagrams which contribute to $`R_{n,m}(f_1\mathrm{}f_n;g_1\mathrm{}g_m)`$ have the property that each $`f_j`$-vertex is connected with at least one $`g_k`$-vertex. (ii) $`R_{n,m}(f_1\mathrm{}f_n;g_1\mathrm{}g_m)=𝒪(\mathrm{}^n)`$. Proof: (i) We work with the notation $`R_{n,m}(Y;X),Y\{y_1,\mathrm{},y_n\},X\{x_1,\mathrm{},x_m\}`$ (cf. ), and consider a subdiagram with vertices $`JY`$ which is not connected with the other vertices $`(YJ)X`$. Because disconneted diagrams factorize with respect to the classical product (58), the corresponding contribution to $`R_{n,m}(Y;X)`$ (55) reads $$\underset{IY}{}(1)^{|I|}\left(\overline{T}(IJ^c)\overline{T}(IJ)\right)\times _{\mathrm{}}\left(T(I^cJ)T(I^cJ^c,X)\right).$$ (65) However, this expression vanishes due to $`_{PJ}(1)^{|P|}\overline{T}(P)\times _{\mathrm{}}T(JP)=0`$ (the latter equation is equivalent to (13), it is the perturbative version of $`S^1S=\mathrm{𝟏}`$). Hence for non-vanishing diagrams $`J`$ must be the empty set. (ii) We express the $`R`$-product in terms of the connected $`T`$\- and $`\overline{T}`$-products $`R_{n,m}(f_1\mathrm{}f_n;g_1\mathrm{}g_m)={\displaystyle \underset{I\{1,\mathrm{},n\}}{}}(1)^{|I|}{\displaystyle \underset{P\mathrm{Part}(I)}{}}{\displaystyle \underset{Q\mathrm{Part}(I^c\{1,\mathrm{},m\})}{}}`$ $`\left({\displaystyle \underset{pP}{}}\overline{T}_{|p|}^c(_{ip}f_i)\right)\times _{\mathrm{}}\left({\displaystyle \underset{qQ}{}}T_{|q|}^c(_{iq}f_i_{jq}g_j)\right)`$ (66) where again $``$ means the classical product (58) and $``$ stands again for the disjoint union. From (63) we know $$\underset{pP}{}\overline{T}_{|p|}^c(_{ip}f_i)=𝒪(\mathrm{}^{|I||P|}),\underset{qQ}{}T_{|q|}^c(_{iq}f_i_{jq}g_j)=𝒪(\mathrm{}^{|I^c|+m|Q|}).$$ (67) From part (i) we conclude that the terms of lowest order (in $`\mathrm{}`$) in $$\left(\underset{pP}{}\overline{T}_{|p|}^c(\mathrm{})\right)\times _{\mathrm{}}\left(\underset{qQ}{}T_{|q|}^c(\mathrm{})\right)=\underset{n0}{}\mathrm{}^nM_n(\underset{pP}{}\overline{T}_{|p|}^c(\mathrm{}),\underset{qQ}{}T_{|q|}^c(\mathrm{}))$$ (68) do not contribute. For simplicity we first consider the special case $`m=1`$. Then only connected diagrams contribute. Hence we obtain $`n|P|+|Q|1`$ similarly to the reasoning after (64). For arbitrary $`m1`$ the terms with minimal power in $`\mathrm{}`$ correspond to diagrams which are maximally disconnected. According to part (i) these diagrams have $`m`$ disconnected components each component containing precisely one vertex $`g_j`$. Applying the $`m=1`$-argument to each of this components we get $`n|P|+|Q|m`$. Taking (67) into account it results the assertion: $`(|I||P|)+(|I^c|+m|Q|)+(|P|+|Q|m)=n`$. ### 5.3 Interacting fields We first describe the perturbative construction of the interacting classical field. Let $``$ be a function of the field which serves as the interaction Lagrangian (for simplicity, we do not consider derivative couplings). We want to find a Poisson algebra generated by a solution of the field equation $$(\text{ }\text{ }\text{ }\text{ }+m^2)\phi _{}(x)=\left(\frac{}{\phi }\right)_{}(x),$$ (69) with the initial conditions $$\begin{array}{cc}\hfill \{\phi _{}(0,𝐱),\phi _{}(0,𝐲)\}=& 0=\{\dot{\phi }_{}(0,𝐱),\dot{\phi }_{}(0,𝐲)\}\hfill \\ \hfill \{\phi _{}(0,𝐱),\dot{\phi }_{}(0,𝐲)\}& =\delta (𝐱𝐲).\hfill \end{array}$$ (70) We proceed in analogy to the construction of the interacting quantum field in Sect. 3 and construct in a first step solutions with localized interactions $`\theta `$ with $`\theta 𝒟(^4)`$ which coincide at early times with the free field (hence the initial conditions (70) are trivially satisfied for sufficiently early times). They are given by a formal power series in the Poisson algebra of the free field $`\phi _\theta (x)={\displaystyle \underset{n=0}{\overset{\mathrm{}}{}}}{\displaystyle _{y_1^0y_2^0\mathrm{}y_n^0x^0}}𝑑y_1𝑑y_2\mathrm{}𝑑y_n\theta (y_1)\mathrm{}\theta (y_n)`$ $`\{(y_1),\{(y_2),\mathrm{}\{(y_n),\phi (x)\}\mathrm{}\}\}`$ (71) Analogous to the quantum case, the structure of the Poisson algebra associated to a causally closed region $`𝒪`$ does not depend on the behaviour of the interaction Lagrangian outside of $`𝒪`$, i.e. there is, for $`\theta ,\theta ^{}\mathrm{\Theta }(𝒪)`$ a canonical transformation $`v`$ with $`v(\phi _\theta (x))=\phi _\theta ^{}(x)`$ for all $`x𝒪`$. The interacting field $`\phi _{}`$ may then be defined as a covariantly constant section within a bundle of Poisson algebras. Starting from the classical interacting field, one may try to define the quantized interacting field by replacing products of free classical fields by the normally ordered product of the corresponding free quantum fields (as in sect. 5.1) and the Poisson brackets in (71) by commutators $$\{,\}\frac{1}{i\mathrm{}}[,]_{\mathrm{}}$$ (72) where the commutator refers to the quantized product $`\times _{\mathrm{}}`$. Note that in general this replacement produces additional terms, e.g. the terms $`k2`$ in $`{\displaystyle \frac{1}{i\mathrm{}}}[:\phi ^n(x):`$ , $`:\phi ^m(y):]_{\mathrm{}}={\displaystyle }_{k=1}^{\mathrm{min}\{n,m\}}(i\mathrm{})^{k1}{\displaystyle \frac{n!m!}{(nk)!(mk)!}}`$ (73) $`\left(\mathrm{\Delta }_+(xy)^k\mathrm{\Delta }_+(yx)^k\right):\phi ^{(nk)}(x)\phi ^{(mk)}(y):`$ which correspond to loop diagrams. Due to the distributional character of the fields with respect to the quantized product the integral in (71), as it stands, is not well defined (there is an ambiguity for coinciding points due to the time ordering). But as we will see Bogoliubov’s formula (29) for the interacting quantum field as a functional derivative of the relative $`S`$-matrix may be interpreted as a precise version of this integral. From the factorization property (11), (14) of time ordered and anti-time ordered products, one gets the following recursion formula for the retarded products (54-55): if $`\mathrm{supp}g`$ is contained in the past and $`\mathrm{supp}f,\mathrm{supp}h`$ in the future of some Cauchy surface, we find $$R_{n+1,m}(gh^n;f^m)=[T_1(g),R_{n,m}(h^n;f^m)]_{\mathrm{}}$$ (74) where we used the fact that $`\overline{T}_1=T_1`$. Hence, for $`m=1`$ and $`y_iy_jij`$ the retarded product $`R_{n,1}(y_1,\mathrm{},y_n;x)`$ can be written in the form<sup>7</sup><sup>7</sup>7The notation for the time ordered products introduced in section 2 is used here for the retarded products. $`R((y_1)\mathrm{}(y_n);\phi (x))=(1)^n{\displaystyle \underset{\pi 𝒮_n}{}}\mathrm{\Theta }(x^0y_{\pi n}^0)\mathrm{\Theta }(y_{\pi n}^0y_{\pi (n1)}^0)\mathrm{}`$ $`\mathrm{\Theta }(y_{\pi 2}^0y_{\pi 1}^0)[(y_{\pi 1}),[(y_{\pi 2})\mathrm{}[(y_{\pi n}),\phi (x)]_{\mathrm{}}\mathrm{}]_{\mathrm{}}]_{\mathrm{}}.`$ (75) (Due to the locality of the interaction $``$ this is a Poincaré covariant expression.) This formula confirms part (ii) of Proposition 2 for non-coinciding $`y_i`$. Our main application of (75) is the study of the classical limit $`\mathrm{}0`$ of the quantized interacting field (29). Due to Proposition 2 (part (ii)) $`R(\mathrm{}^1(y_1)\mathrm{}\mathrm{}^1(y_n);\phi (x))`$ contains no terms with negative powers of $`\mathrm{}`$ and thus has a well-defined classical limit. We conclude that the quantized interacting field (29), (54) $$\phi _\theta (h)=\underset{n=0}{\overset{\mathrm{}}{}}\frac{i^n}{n!\mathrm{}^n}R_{n,1}((\theta )^n;h\phi ),h𝒟(^4),$$ (76) tends to the classical interacting field (71) in this limit. Note that the factor $`\mathrm{}^1`$ in the interaction Lagrangian is in accordance with the quantization rule (72), since in (75) there is for each factor $``$ precisely one commutator. In $`R_{n,1}((\theta )^n;f\phi )`$ the above mentioned ambiguities for coinciding points in the iterated retarded commutators have been fixed by the definition of time ordered products as everywhere defined distributions. The normalization condition N4 implies an analogous equation for the retarded product $`R_{n,1}`$ (cf. ). The latter means that $`\phi _{}`$ (76) satisfies the same field equation as the classical interacting field (69) $$(\text{ }\text{ }\text{ }\text{ }+m^2)\phi _{}(x)=\left(\frac{}{\phi }\right)_{}(x).$$ (77) Here $`\left(\frac{}{\phi }\right)_{}`$ is not necessarily a polynomial in $`\phi _{}`$ (the pointwise product of interacting fields is in general not defined). We found that the relative $`S`$-matrices $`S_{\mathrm{}^1\theta }(f)(f𝒟(^4,𝒱))`$, and hence all elements of the algebra $`𝒜_{\mathrm{}^1\theta }`$ are power series in $`\mathrm{}`$. For the global algebras of covariantly constant sections we recall from that the unitaries $`V𝒰(\theta ,\theta ^{})`$ can be chosen as relative $`S`$-matrices $$V=S_{\mathrm{}^1\theta }(\mathrm{}^1\theta _{})^1𝒰(\theta ,\theta ^{})$$ (78) where $`\theta _{}𝒟(^4)`$ depends in the following way on $`(\theta \theta ^{})`$: we split $`\theta \theta ^{}=\theta _++\theta _{}`$ with $`\mathrm{supp}\theta _+(C(𝒪)+\overline{V}_{})=\mathrm{}`$ and $`\mathrm{supp}\theta _{}(C(𝒪)+\overline{V}_+)=\mathrm{}`$ (where $`C(𝒪)`$ means the causally closed region containing $`𝒪`$ in which $`\theta `$ and $`\theta ^{}`$ agree, cf. (18)). So $`V`$ is a formal Laurent series in $`\mathrm{}`$, and the sections are no longer well defined power series. Replacing $`𝒜`$ and $`𝒜(𝒪)`$ by $`_{n_0}\mathrm{}^n𝒜`$ and $`_{n_0}\mathrm{}^n𝒜(𝒪)`$ (for the new algebras the same symbol $`𝒜`$ will be used again) we obtain modules over the ring of formal power series in $`\mathrm{}`$ with complex coefficients. For the further construction the validity of part (iii) of the following Proposition is crucial: Proposition 3: (i) Let $`R_{n,m}(\mathrm{};\mathrm{})=_{a=1}^mR_{n,m}^{(a)}(\mathrm{};\mathrm{})`$ where $`R_{n,m}^{(a)}(\mathrm{};\mathrm{})`$ is the sum of all diagrams with $`a`$ connected components. Then $$R_{n,m}^{(a)}((\mathrm{}^1\theta )^n;(\mathrm{}^1\theta _{})^m)=𝒪(\mathrm{}^a).$$ (79) (Note that the range of $`a`$ is restricted by part (i) of Proposition 2.) This estimate is of more general validity: instead of a retarded product we could have e.g. a multiple $`\times _{\mathrm{}}`$-product, a time ordered or antichronological product and the factors may be quite arbitrary. It is only essential that each factor is of order $`𝒪(\mathrm{}^1)`$. (ii) Let $`A𝒜(𝒪)`$. Then all diagrams which contribute to $`V\times _{\mathrm{}}A\times _{\mathrm{}}V^1`$ (where $`V`$ is given by (78)) have the property that each vertex of $`V`$ and of $`V^1`$ is connected with at least one vertex of $`A`$. (It may happen that a connected component of $`V`$ is not directly connected with $`A`$, but that it is connectecd with a connected component of $`V^1`$ and the latter is connected with $`A`$.) (iii) $$𝒜(𝒪)A=𝒪(\mathrm{}^n)V\times _{\mathrm{}}A\times _{\mathrm{}}V^1=𝒪(\mathrm{}^n)$$ (80) In particular if $`A`$ is the term of $`n`$-th order in $`\mathrm{}`$ of an interacting field, then $`V\times _{\mathrm{}}A\times _{\mathrm{}}V^1`$ is a power series in $`\mathrm{}`$ in which the terms up to order $`\mathrm{}^{n1}`$ vanish. Proof: Part (i) is obtained essentially in the same way as Proposition 1. Part (iii) is a consequence of parts (i) and (ii), and the following observation: let us consider a diagram which contributes to $`V\times _{\mathrm{}}A\times _{\mathrm{}}V^1`$ according to part (ii). If the subdiagrams belonging to $`V`$ and $`V^1`$ have $`r`$ and $`s`$ connected components, then the whole diagram has at least $`(r+s)`$ contractions, which yield a factor $`\mathrm{}^{(r+s)}`$. It remains the proof of (ii): We use the same notations as in the proof of Proposition 2. Let $`Y_1Y_2=Y`$, $`X_1X_2=X`$. We now consider the sum of all diagrams contributing to $`R(Y,X)`$ in which the vertices $`(Y_1,X_1)`$ are not connected with the vertices $`(Y_2,X_2)`$. Using (55) and the fact that disconnected diagrams factorize with respect to the classical product (58), this (partial) sum is equal to $`{\displaystyle \underset{IY}{}}(1)^{|IY_1|}[\overline{T}(IY_1)\times _{\mathrm{}}T(I^cY_1,X_1)]`$ $`(1)^{|IY_2|}[\overline{T}(IY_2)\times _{\mathrm{}}T(I^cY_2,X_2)]`$ $`=R(Y_1,X_1)R(Y_2,X_2).`$ (81) From $`\mathrm{𝟏}=VV^1=VV^{}`$, (54) and (78) we know $$\underset{Y_1Y_2=Y,X_1X_2=X}{}(1)^{(|Y_1|+|X_1|)}R^{}(Y_1,X_1)\times _{\mathrm{}}R(Y_2,X_2)=0$$ (82) for fixed $`(Y,X)`$, $`YX\mathrm{}`$. Next we note $`V\times _{\mathrm{}}A\times _{\mathrm{}}V^1={\displaystyle \underset{n,m}{}}{\displaystyle \frac{1}{n!m!}}{\displaystyle 𝑑y_1\mathrm{}𝑑y_n𝑑x_1\mathrm{}𝑑x_m\theta (y_1)\mathrm{}\theta (y_n)\theta _{}(x_1)\mathrm{}\theta _{}(x_m)}`$ $`{\displaystyle \underset{Y_1Y_2=Y,X_1X_2=X}{}}(i)^{(|Y_1|+|X_1|)}i^{(|Y_2|+|X_2|)}R^{}(Y_1,X_1)\times _{\mathrm{}}A\times _{\mathrm{}}R(Y_2,X_2),`$ (83) where we have used the notations $`Y\{y_1,\mathrm{}y_n\},X\{x_1,\mathrm{},x_n\}`$. In the integrand of the latter expression we consider (for given $`Y`$ and $`X`$) fixed decompositions $`Y=Y_3Y_4`$ and $`X=X_3X_4`$, $`Y_3X_3\mathrm{}`$. Now we consider the (partial) sum of all diagrams in which the vertices $`(Y_3,X_3)`$ are not connected with $`A`$ and each of the vertices $`(Y_4,X_4)`$ is connected with $`A`$. Part (ii) is proved if we can show that this partial sum vanishes. This holds in fact true because $`R^{}`$ and $`R`$ factorize according to (81), and due to the unitarity (82): $`{\displaystyle \underset{Y_1Y_2=Y,X_1X_2=X}{}}(1)^{(|Y_1Y_4|+|X_1X_4|)}[R^{}(Y_1Y_4,X_1X_4)\times _{\mathrm{}}`$ $`A\times _{\mathrm{}}R(Y_2Y_4,X_2X_4)]`$ $`(1)^{(|Y_1Y_3|+|X_1X_3|)}[R^{}(Y_1Y_3,X_1X_3)\times _{\mathrm{}}R(Y_2Y_3,X_2X_3)]=0.\text{ }\text{ }\text{ }`$ Now we are ready to give an algebraic formulation of the expansion in $`\mathrm{}`$. Let $`I_n\stackrel{\mathrm{def}}{=}\mathrm{}^n𝒜_{}`$. $`I_n`$ is an ideal in the global algebra $`𝒜_{}`$. We define $$𝒜_{}^{(n)}\stackrel{\mathrm{def}}{=}\frac{𝒜_{}}{I_{n+1}},𝒜_{}^{(n)}(𝒪)\stackrel{\mathrm{def}}{=}\frac{𝒜_{}(𝒪)}{I_{n+1}𝒜_{}(𝒪)}.$$ (84) which means that we neglect all terms which are of order $`𝒪(\mathrm{}^{n+1})`$. The embeddings $`i_{21}:𝒜_{}(𝒪_1)𝒜_{}(𝒪_2)`$ for $`𝒪_1𝒪_2`$ induce embeddings $`𝒜_{}^{(n)}(𝒪_1)𝒜_{}^{(n)}(𝒪_2)`$. Thus we obtain a projective system of local nets $`(𝒜_{}^{(n)}(𝒪))`$ of algebras of quantum observables up to order $`\mathrm{}^{n+1}`$. Note that we may equip our algebras $`𝒜_{}^{(n)}`$ also with the Poisson bracket induced by $`\frac{1}{i\mathrm{}}[,]_{\mathrm{}}`$, because the ideals $`I_n`$ are also Poisson ideals with respect to these brackets. Then $`𝒜_{}^{(0)}`$ becomes the local net of Poisson algebras of the classical field theory, whereas for $`n0`$ we obtain a net of noncommutative Poisson algebras. The expansion in powers of $`\mathrm{}`$ is usually called “loop expansion”. This is due to the fact that the order in $`\mathrm{}`$ of a certain Feynman diagram belonging to $`R_{n,m}((\mathrm{}^1\theta )^n;f_1\mathrm{}f_m),𝒟(^4,𝒱)f_j=𝒪(\mathrm{}^0)`$, is equal to: (number of propagators (i.e. inner lines)) - $`n`$ $`=`$ (number of loops) + $`m`$ \- (number of connected components). In particular, using part (i) of Proposition 2, we find that for the interacting fields ($`m=1`$) the order in $`\mathrm{}`$ agrees with the number of loops. ## 6 Local algebraic formulation of the quantum action principle The method of algebraic renormalization (for an overview see ) relies on the so called ’quantum action principle’ (QAP), which is due to Lowenstein and Lam . This principle is a formula for the variation of (possibly connected or one-particle-irreducible) Green’s functions (or of the corresponding generating functional) under \- a change of coordinates (e.g. one applies the differential operator of the field equation to the Green’s functions), \- a variation of the fields (e.g. the BRST-transformation) \- a variation of a parameter. This may be a parameter in the Lagrangian or in the normalization conditions for the Green’s functions. These are different theorems with different proofs. The common statement is that the variation of the Green’s functions is equal to the insertion of a local or spacetime integrated composite field operator (for details see ). In this section we study two simple cases of the QAP: the field equation and the variation of a parameter which appears only in the interaction Lagrangian. The aim of this section is to formulate the QAP (in these two cases) for our local algebras of observables $`𝒢_{}(𝒪)`$, i.e. we are looking for an operator identity which holds true independently of the adiabatic limit. Such an identity does not depend on the choice of a state, as it is the case for the Green’s functions. In a second step we compare our formula with the usual formulation of the QAP in terms of Green’s functions. The latter are the vacuum expectation values in the adiabatic limit $`g1`$.<sup>8</sup><sup>8</sup>8This limit is taken by scaling the test function $`g`$: let $`g_0𝒟(^4),g_0(0)=1`$; then one considers the limit $`ϵ0(ϵ>0)`$ of $`g_ϵ(x)g_0(ϵx)`$. Uniqueness of the adiabatic limit means the independence of the particular choice of $`g_0`$. We specialize to models for which the adiabatic limit is known to exist. This is the case for pure massive theories and certain theories with (some) massless particles such as QED and $`\lambda :\phi ^{2n}:`$-theories , provided the time ordered products are appropriately normalized. Remarks: (1) From the usual QAP (in terms of Green’s functions) one obtains an operator identity by means of the Lehmann-Symanzik-Zimmermann - reduction formalism . Although the latter relies on the adiabatic limit an analogous conclusion from the Fock vacuum expectation values to arbitrary matrix elements is possible in our local construction: let $`𝒪`$ be an open double cone and let $`x_1,\mathrm{},x_k((\overline{𝒪}\{x_{k+l+1},\mathrm{},x_n\})+\overline{V}_{}),x_{k+1},\mathrm{},x_{k+l}𝒪`$ and $`x_{k+l+1},\mathrm{},x_n(\overline{𝒪}+\overline{V}_+)`$. Using the causal factorization of time ordered products of interacting fields (28) we obtain $`(\mathrm{\Omega },T_\theta \left(\phi (x_1)\mathrm{}\phi (x_n)\right)\mathrm{\Omega })=(T_\theta \left(\phi (x_1)\mathrm{}\phi (x_k)\right)^{}\mathrm{\Omega },`$ $`T_\theta \left(\phi (x_{k+1})\mathrm{}\phi (x_{k+l})\right)T_\theta \left(\phi (x_{k+l+1})\mathrm{}\phi (x_n)\right)\mathrm{\Omega }).`$ (85) Now we choose $`\theta \mathrm{\Theta }(𝒪)`$ such that $`\{x_1,\mathrm{},x_k\}(\mathrm{supp}\theta +\overline{V}_{})=\mathrm{}`$ and $`\{x_{k+l+1},\mathrm{},x_n\}(\mathrm{supp}\theta +\overline{V}_+)=\mathrm{}`$. Due to the retarded support (56) of the $`R`$-products we then know that $`T_\theta \left(\phi (x_{k+l+1})\mathrm{}\phi (x_n)\right)`$ agrees with the time ordered product $`T_0\left(\phi (x_{k+l+1})\mathrm{}\phi (x_n)\right)`$ of the corresponding free fields. By means of $`S_\theta (f\phi )=S(\theta )^1S(f\phi )S(\theta )`$ for $`\mathrm{supp}f(\mathrm{supp}\theta +\overline{V}_{})=\mathrm{}`$ we obtain $$T_\theta \left(\phi (x_1)\mathrm{}\phi (x_k)\right)^{}=S(\theta )^1T_0\left(\phi (x_1)\mathrm{}\phi (x_k)\right)^{}S(\theta ).$$ (86) Our assertion follows now from the fact that the states $`T_0\left(\phi (x_{k+l+1})\mathrm{}\phi (x_n)\right)\mathrm{\Omega }`$ generate a dense subspace of the Fock space and the same for the states $`S(\theta )^1T_0\left(\phi (x_1)\mathrm{}\phi (x_k)\right)^{}S(\theta )\mathrm{\Omega }`$. (For the validity of the latter statement it is important that $`x_1,\mathrm{},x_k`$ can be arbitrarily spread over a Cauchy surface which is later than $`(\overline{𝒪}\{x_{k+l+1},\mathrm{},x_n\})`$.) (2) Recently Pinter presented an alternative derivation of the QAP for the variation of a parameter in the Lagrangian (including the free part) also in the framework of causal perturbation theory. In contrast to our presentation Pinter’s QAP is formulated for the $`S`$-matrix and in the adiabatic limit. The main new technical tool which is used is a generalization of the normalization condition N4. ### 6.1 Field equation The normalization condition N4 implies $`(\text{ }\text{ }\text{ }_x+m^2)R((y_1)\mathrm{}(y_n);\varphi (x)\varphi (x_1)\mathrm{}\varphi (x_m))=`$ $`i{\displaystyle \underset{l=1}{\overset{n}{}}}\delta (xy_l)R((y_1)\mathrm{}\widehat{l}\mathrm{}(y_n);{\displaystyle \frac{}{\varphi }}(x)\varphi (x_1)\mathrm{}\varphi (x_m))`$ $`i{\displaystyle \underset{j=1}{\overset{m}{}}}\delta (xx_j)R((y_1)\mathrm{}(y_n);\varphi (x_1)\mathrm{}\widehat{j}\mathrm{}\varphi (x_m)),`$ (87) where $`\widehat{l}`$ and $`\widehat{j}`$ means that the corresponding factor is omitted. This equation takes a simple form for the corresponding generating functionals (i.e. the relative $`S`$-matrices (16)) $$f(x)S_g(f\varphi )=(\text{ }\text{ }\text{ }\text{ }_x+m^2)\frac{\delta }{i\delta f(x)}S_g(f\varphi )\frac{\delta }{i\delta \rho (x)}|_{\rho =0}S_g(f\varphi +\rho g\frac{}{\varphi }).$$ (88) To formulate this in terms of our local algebras of observables (cf. sect. 3) we set $`g\theta \mathrm{\Theta }(𝒪)`$ and for $`x𝒪`$ we can choose $`\rho `$ such that $`\mathrm{supp}\rho \{y|\theta (y)=1\}`$. Then (88) turns into $$(\text{ }\text{ }\text{ }\text{ }_x+m^2)\frac{\delta }{i\delta f(x)}S_{}(f\varphi )=f(x)S_{}(f\varphi )+\frac{\delta }{i\delta \rho (x)}|_{\rho =0}S_{}(f\varphi +\rho \frac{}{\varphi }),x𝒪.$$ (89) This is the QAP (in the case of the field equation) for the local algebras of observables. To compare with the usual form of the QAP we consider the generating functional $`Z(f)`$ for the Green’s functions $`<\mathrm{\Omega }|T\left(\varphi _{}(x_1)\mathrm{}\varphi _{}(x_m)\right)|\mathrm{\Omega }>`$ which is obtained from the relative $`S`$-matrices by $$Z(f)=\underset{g1}{lim}(\mathrm{\Omega },S_g(f\varphi )\mathrm{\Omega }),$$ (90) where $`\mathrm{\Omega }`$ is the Fock vacuum . So by taking the vacuum expectation value and the adiabatic limit of (88) we get $$f(x)Z(f)=\mathrm{\Delta }(x)Z(f),$$ (91) where $`\mathrm{\Delta }(x)`$ is a insertion of UV-dimension<sup>9</sup><sup>9</sup>9We assume that $``$ has UV-dimension $`4`$. $`3`$, coinciding with the classical field polynomial $`\frac{\delta S}{\delta \varphi (x)}`$ in the classical approximation (where $`S=d^4x[\frac{1}{2}(_\mu \varphi (x)^\mu \varphi (x)m^2\varphi ^2(x))+g(x)(x)]`$ is the classical action). Equation (91) is the usual form of the QAP (cf. eqn. (3.20) in ). In the present case the local algebraic formulation (89) contains more information than the usual QAP (91). ### 6.2 Variation of a parameter in the interaction In (54) we have defined retarded products of Wick polynomials, i.e. elements of the Borchers class. Analogously we now introduce retarded products $`R_{}(g^n;f^m)`$ of interacting fields $$S_{+g}(f)=S_{}(g)^1S_{}(g+f)\stackrel{\mathrm{def}}{=}\underset{n,m=0}{\overset{\mathrm{}}{}}\frac{i^{n+m}}{n!m!}R_{}(g^n;f^m)$$ (92) where $`,g,f𝒟(^4,𝒱)`$. Obviously they can be expressed in terms of antichronological and time ordered products of interacting fields by exactly the same formula as in the case of Wick polynomials (55) $$R_{}(g^n;f^m)=\underset{k=0}{\overset{n}{}}(1)^k\frac{n!}{k!(nk)!}\overline{T}_{}(g^k)T_{}(g^{(nk)}f^m).$$ (93) Thereby the antichronological product of interacting fields is defined analogously to the time ordered product (28), namely by $$\overline{T}_{}(f^m)=\frac{d^m}{(i)^md\lambda ^m}|_{\lambda =0}S_{}(\lambda f)^1,$$ (94) and satisfies anticausal factorization (14) (which justifies the name). The support property (56) of the retarded products relies on the (anti)causal factorization of the $`T`$\- and $`\overline{T}`$-products (11, 14), hence, the $`R`$-product of interacting fields (92-93) has also retarded support (56). Similarly to Lowenstein in sect.II.B we consider an infinitesimal change of the interaction Lagrangian $$_0_0+ϵ_1$$ (95) where $`_0,_1𝒱`$ or $`𝒟(^4,𝒱)`$. For the $`m`$-fold variation of the time ordered product of the interacting fields (28) we obtain $`{\displaystyle \frac{d^m}{dϵ^m}}|_{ϵ=0}T_{\theta (_0+ϵ_1)}(f^l)`$ $`=`$ $`{\displaystyle \frac{^m}{ϵ^m}}|_{ϵ=0}{\displaystyle \frac{^l}{i^l\lambda ^l}}|_{\lambda =0}S_{\theta (_0+ϵ_1)}(\lambda f)`$ (96) $`=`$ $`i^mR_{\theta _0}((\theta _1)^m;f^l).`$ To formulate this identity for our local algebras of observables we assume that the $`_1`$ has compact support, i.e. $`_1𝒟(^4,𝒱)`$. We set $$\mathrm{\Theta }_0(𝒪)\stackrel{\mathrm{def}}{=}\{\theta \mathrm{\Theta }(𝒪)|\theta |_{\mathrm{supp}_1}1\}.$$ (97) We consider the observables as covariantly constant sections in the bundle over $`\mathrm{\Theta }_0(𝒪)`$ (instead of $`\mathrm{\Theta }(𝒪)`$ as in sect. 3). Then we obtain $$\frac{d^m}{dϵ^m}|_{ϵ=0}T_{_0+ϵ_1}(f^l)=i^mR__0(_1^m;f^l).$$ (98) This is the local algebraic formulation of the QAP for the variation of a parameter in the interaction. We are now going to investigate the usual QAP by using Epstein and Glaser’s definition of Green’s functions (90). In (96) the $`m`$-fold variation of the parameter $`ϵ`$ results in a retarded insertion of $`(\theta _1)^m`$. In the usual QAP $`(\theta _1)^m`$ is inserted into the time ordered product, i.e. one considers $$i^mT_{\theta _0}((\theta _1)^mf^l)=\frac{^m}{ϵ^m}|_{ϵ=0}\frac{^l}{i^l\lambda ^l}|_{\lambda =0}S_{\theta _0}(\theta ϵ_1+\lambda f).$$ (99) Obviously (96) and (99) do not agree. However, let us assume that we are dealing with a pure massive theory and that $`_0`$ and $`_1`$ have UV-dimension $`\mathrm{dim}(_j)=4`$. Or: if $`\mathrm{dim}(_j)<4`$ we assume that $`_j`$ is treated in the extension to the total diagonal as if it would hold $`\mathrm{dim}(_j)=4`$. Hence it may occur that the scaling degree increases in the extension to a certain amount: $`\mathrm{sd}(t_0)\mathrm{sd}(t)4nb`$ for a scalar theory without derivative couplings, where $`b`$ is the number of external legs (cf. (51)-(53)). (In the BPHZ framework one says that $`_j`$ is ’oversubtracted with degree $`4`$’.) Then there exists a normalization of the time ordered products, which is compatible with the other normalization conditions N1 \- N4 and (63), such that the Green’s functions corresponding to (99) exist and agree, i.e. we assert $$\frac{d^m}{dϵ^m}|_{ϵ=0}\underset{\theta 1}{lim}(\mathrm{\Omega },T_{\theta (_0+ϵ_1)}(f^l)\mathrm{\Omega })=i^m\underset{\theta 1}{lim}(\mathrm{\Omega },T_{\theta _0}((\theta _1)^mf^l)\mathrm{\Omega })$$ (100) for all $`m,l_0`$, which is equivalent to $$\underset{\theta 1}{lim}(\mathrm{\Omega },S_{\theta (_0+ϵ_1)}(\lambda f)\mathrm{\Omega })=\underset{\theta 1}{lim}(\mathrm{\Omega },S_{\theta _0}(\theta ϵ_1+\lambda f)\mathrm{\Omega }).$$ (101) (We assume that the derivatives $`\frac{^m}{ϵ^m}`$ and $`\frac{^l}{\lambda ^l}`$ commute with the adiabatic limit $`\theta 1`$. This seems to be satisfied for vacuum expectation values in pure massive theories as it is the case here .) This is the usual form of the QAP (in terms of Epstein and Glaser’s Green’s functions) for the present case (cf. eqn. (2.6) of <sup>10</sup><sup>10</sup>10Lowenstein works with Zimermanns definition of normal products of interacting fields: $`N_\delta \{_{j=1}^l\phi _{i_j}(x)\},\delta d_{j=1}^ld(\phi _{i_j})`$ . For $`\delta =d`$ (i.e. without oversubtraction) $`N_\delta \{_{j=1}^l\phi _{i_j}(x)\}`$ agrees essentially with our $`(:_{j=1}^l\phi _{i_j}(x):)_g`$ (29). The difference is due to the adiabatic limit and the different ways of defining Green’s functions (Zimmermann uses the Gell-Mann Low series, cf. (102), (106)).). In contrast to the field equation, the QAP (100) does not hold for the operators before the adiabatic limit. Proof of (100): For a better comparison with Lowenstein’s formulation, we present a proof which makes the detour over the corresponding Gell-Mann Low expressions. First we comment on the equality of Epstein and Glaser’s Green’s functions with the Gell-Mann Low series $$\underset{\theta 1}{lim}(\mathrm{\Omega },S_\theta (f)\mathrm{\Omega })=\underset{\theta 1}{lim}\frac{(\mathrm{\Omega },S(\theta +f)\mathrm{\Omega })}{(\mathrm{\Omega },S(\theta )\mathrm{\Omega })},$$ (102) which is proved in the appendix of . This can be understood in the following way: let $`P_\mathrm{\Omega }`$ be the projector on the Fock vacuum $`\mathrm{\Omega }`$ and $`P_\mathrm{\Omega }^{}\stackrel{\mathrm{def}}{=}\mathrm{\hspace{0.17em}1}P_\mathrm{\Omega }`$. Using $`S(\theta )^{}=S(\theta )^1`$ we obtain $`(\mathrm{\Omega },S_\theta (f)\mathrm{\Omega })`$ $`=`$ $`(S(\theta )\mathrm{\Omega },(P_\mathrm{\Omega }+P_\mathrm{\Omega }^{})S(\theta +f)\mathrm{\Omega })`$ (103) $`=`$ $`{\displaystyle \frac{(\mathrm{\Omega },S(\theta +f)\mathrm{\Omega })}{(\mathrm{\Omega },S(\theta )\mathrm{\Omega })}}|(\mathrm{\Omega },S(\theta )\mathrm{\Omega })|^2`$ $`+`$ $`(\mathrm{\Omega },S(\theta )^1P_\mathrm{\Omega }^{}S(\theta +f)\mathrm{\Omega })`$ and $`1`$ $`=`$ $`(\mathrm{\Omega },S(\theta )^1(P_\mathrm{\Omega }+P_\mathrm{\Omega }^{})S(\theta )\mathrm{\Omega })`$ (104) $`=`$ $`|(\mathrm{\Omega },S(\theta )\mathrm{\Omega })|^2+(\mathrm{\Omega },S(\theta )^1P_\mathrm{\Omega }^{}S(\theta )\mathrm{\Omega }).`$ In $`(\mathrm{\Omega },S(\theta )^1P_\mathrm{\Omega }^{}S(\theta +f)\mathrm{\Omega })`$ there is at least one contraction between $`S(\theta )^1`$ and $`S(\theta +f)`$ (or: the terms without contraction are precisely $`(\mathrm{\Omega },S(\theta )^1\mathrm{\Omega })(\mathrm{\Omega },S(\theta +f)\mathrm{\Omega })`$). In the mentioned reference the support properties in momentum space of the contracted terms are analysed and in this way it is proved $$\underset{\theta 1}{lim}(\mathrm{\Omega },S(\theta )^1P_\mathrm{\Omega }^{}S(\theta +f)\mathrm{\Omega })=0.$$ (105) Inserting this into (103) and (with $`f=0`$) into (104) it results (102). Because of (102) our assertion (101) is equivalent to $$\underset{\theta 1}{lim}\frac{(\mathrm{\Omega },S(\theta (_0+ϵ_1)+\lambda f)\mathrm{\Omega })}{(\mathrm{\Omega },S(\theta (_0+ϵ_1))\mathrm{\Omega })}=\underset{\theta 1}{lim}\frac{(\mathrm{\Omega },S(\theta (_0+ϵ_1)+\lambda f)\mathrm{\Omega })}{(\mathrm{\Omega },S(\theta _0)\mathrm{\Omega })}.$$ (106) This is the QAP in terms of the Gell-Mann Low series. Obviously the nontrivial statement is $$\underset{\theta 1}{lim}\frac{(\mathrm{\Omega },S(\theta (_0+ϵ_1))\mathrm{\Omega })}{(\mathrm{\Omega },S(\theta _0)\mathrm{\Omega })}=1.$$ (107) A possibility to ensure the validity of this equation is the above assumption (which has not been used so far) that $`_0`$ and $`_1`$ have mass dimension $`\mathrm{dim}(_j)4`$ and are treated as dimension $`4`$ vertices in the renormalization procedure. Due to this additional assumption and the requirements that the adiabatic limit exists and is unique, the normalization of the vacuum diagrams is uniquely fixed, and with this normalization the vacuum diagrams vanish in the adiabatic limit $$\underset{\theta 1}{lim}(\mathrm{\Omega },S(\theta _0)\mathrm{\Omega })=1,\underset{\theta 1}{lim}(\mathrm{\Omega },S(\theta (_0+ϵ_1))\mathrm{\Omega })=1.$$ (108) (For a proof see also the appendix of .) Remarks: (1) Without the assumption about $`_0`$ and $`_1`$ we find $$\underset{\theta 1}{lim}(\mathrm{\Omega },S_{\theta (_0+ϵ_1)}(\lambda f)\mathrm{\Omega })=\underset{\theta 1}{lim}\frac{(\mathrm{\Omega },S_{\theta _0}(\theta ϵ_1+\lambda f)\mathrm{\Omega })}{(\mathrm{\Omega },S_{\theta _0}(\theta ϵ_1)\mathrm{\Omega })}$$ (109) instead of (101), by using (102) only. This is a formulation of the QAP for general situations in which (107) does not hold. (2) By means of the QAP (98) (or (100), or (109)) one can compute the change of the time ordered products of interacting fields (or of the Green’s functions) under the variation of parameters $`\lambda _1,\mathrm{},\lambda _s`$ if the interaction Lagrangian has the form $`(x)=_ia_i(\lambda _1,\mathrm{},\lambda _s)_i(x),_i𝒱`$ resp. $`𝒟(^4,𝒱)`$ (cf. eqns. (2.7-8) of ). But only the interaction $``$ may depend on the parameters and not the time ordering operator (i.e. the normalization conditions for the time ordered products). ## Appendix: wavefront sets and the pointwise product of distributions In this appendix we briefly recall the definition of the wavefront set of a distribution and mention a simple criterion for the existence of the pointwise product of distributions in terms of their wavefront sets. For a detailed treatment we refer to Hörmander , the application to quantum field theory on curved spacetimes can be found in . Let $`t𝒟^{}(^d)`$ be singular at the point $`x`$ and let $`f𝒟(^4)`$ with $`f(x)0`$. Then $`ft𝒟^{}(^d)`$ is also singular at $`x`$ and $`ft`$ has compact support. Hence the Fourier transform $`\widehat{ft}`$ is a $`𝒞^{\mathrm{}}`$-function. In some directions $`\widehat{ft}`$ does not rapidly decay, because otherwise $`ft`$ would be infinitly differentiable at $`x`$. Thereby a function $`g`$ is called rapidly decaying in the direction $`k^d\{0\}`$, if there is an open cone $`C`$ with $`kC`$ and $`\mathrm{sup}_{k^{}C}|k^{}|^N|g(k^{})|<\mathrm{}`$ for all $`N𝐍`$. Definition: The wavefront set $`\mathrm{WF}(t)`$ of a distribution $`t𝒟^{}(^d)`$ is the set of all pairs $`(x,k)^d\times ^d\{0\}`$ such that the Fourier transform $`\widehat{ft}`$ does not rapidly decay in the direction $`k`$ for all $`f𝒟(^d)`$ with $`f(x)0`$. For example the delta distribution satisfies $`\widehat{f\delta }(k)=f(0)`$, hence $`\mathrm{WF}(\delta )=\{0\}\times ^d\{0\}`$. The wavefront set is a refinement of the singular support of $`t`$ (which is the complement of the largest open set where $`t`$ is smooth): $$t\mathrm{is}\mathrm{singular}\mathrm{at}xk^d\{0\}\mathrm{with}(x,k)\mathrm{WF}(t).$$ For the wavefront set of the two-point function one finds $$\mathrm{WF}(\mathrm{\Delta }_+)=\{(x,k)|x^2=0,k^2=0,xk,k_0>0\}.$$ (110) Let $`t`$ and $`s`$ be two distributions which are singular at the same point $`x`$. We localize them by multiplying with $`f𝒟(^d)`$ where $`f(x)0`$. We assume that $`(ft)`$ and $`(fs)`$ have only one overlapping singularity, namely at $`x`$. In general the pointwise product $`(ft)(y)(fs)(y)`$ does not exist. Heuristically this can be seen by the divergence of the convolution integral $`𝑑k\widehat{(ft)}(pk)\widehat{(fs)}(k)`$. But this integral converges if $`k_1+k_20`$ for all $`k_1,k_2`$ with $`(x,k_1)\mathrm{WF}(t)`$ and $`(x,k_2)\mathrm{WF}(s)`$. This makes plausible the following theorem: Theorem: Let $`t,s𝒟^{}(^d)`$ with $$\{(x,k_1+k_2)|(x,k_1)\mathrm{WF}(t)(x,k_2)\mathrm{WF}(s)\}(^d\times \{0\})=\mathrm{}.$$ (111) Then the pointwise product $`(ts)𝒟^{}(^d)`$ exists. By means of this theorem one verifies the existence of the distributional products $`(\phi ^n)_{\mathrm{}}(t)`$ (44) and $`(t_{k,\mathrm{}}s)`$ (47). Acknowledgements: We thank Gudrun Pinter for several discussions on the quantum action principle, and Volker Schomerus and Stefan Waldmann for discussions on deformation quantization. In particular we are grateful to Stefan Waldmann for drawing our attention to reference .
warning/0001/cond-mat0001408.html
ar5iv
text
# Spanning Trees on Hypercubic Lattices and Non-orientable Surfaces ## 1 INTRODUCTION The problem of enumerating spanning trees on a graph was first considered by Kirchhoff in his analysis of electrical networks. Consider a graph $`G=\{V,E\}`$ consisting of a vertex set $`V`$ and an edge set $`E`$. We shall assume that $`G`$ is connected. A subset of edges $`TE`$ is a spanning tree if it has $`|V|1`$ edges with at least one edge incident at each vertex. Therefore $`T`$ has no cycles. In ensuing discussions we shall use $`T`$ to also denote the spanning tree. Number the vertices from 1 to $`|V|`$ and associate to the edge $`e_{ij}`$ connecting vertices $`i`$ and $`j`$ a weight $`x_{ij}`$, with the convention of $`x_{ii}=0`$. The enumeration of spanning trees concerns with the evaluation of the tree generating function $$T(G;\{x_{ij}\})=\underset{TE}{}\underset{e_{ij}T}{}x_{ij},$$ (1) where the summation is taken over all spanning trees $`T`$. Particularly, the number of spanning trees on $`G`$ is obtained by setting $`x_{ij}=1`$ as $$N_{SPT}(G)=T(G;1).$$ (2) Considerations of spanning tree also arise in statistical physics in the enumeration of close-packed dimers (perfect matchings) . Using a similar consideration, for example, one of us has evaluated the number of spanning trees for the simple quartic, triangular and honeycomb lattices in the limit of $`|V|\mathrm{}`$. In this Letter we report new results on the evaluation of the generating function Eq. (1) for finite hypercubic lattices in arbitrary dimensions. Results are also obtained for a simple quartic net embedded on two non-orientable surfaces, the Möbius strip and the Klein bottle. As the main formula used in this Letter is a relation expressing the tree generating function in terms of the eigenvalues of an associated tree matrix, for completeness we give an elementary derivation of this formula. ## 2 THE TREE MATRIX For a given graph $`G=\{V,E\}`$ consider a $`|V|\times |V|`$ matrix M$`(G)`$ with elements $$M_{ij}(G)=\{\begin{array}{cc}_{k=1}^{|V|}x_{ik},\hfill & i=j=1,2,\mathrm{},|V|\hfill \\ x_{ij},\hfill & \text{if vertices }i\text{}j\text{}ij\text{, are connected by an edge}\hfill \\ 0,\hfill & \text{otherwise.}\hfill \end{array}$$ (3) We shall refer to M$`(G)`$ simply as the tree matrix. It is well-known that the tree generating function, Eq. (1), is given by the cofactor of any element of the tree matrix, and that the cofactor is the same for all elements. Namely, we have the identity $$T(G;\{x_{ij}\})=\text{the cofactor of }\text{any}\text{ element of the matrix }\text{M}\text{(G)}.$$ (4) The tree generating function can also be expressed in terms of the eigenvalues of the tree matrix M$`(G)`$. We give here an elementary derivation of this result which we use in subsequent sections. Let $`𝐌(G)`$ be the tree matrix of a graph $`G=\{V,E\}`$. Since the sum of all elements in a row of $`𝐌(G)`$ equals to zero, $`𝐌(G)`$ has 0 as an eigenvalue and, by definition, we have $$\mathrm{det}\left|M_{ij}(G)\lambda \delta _{ij}\right|=\lambda F(\lambda )$$ (5) where $$F(\lambda )=\underset{i=2}{\overset{|V|}{}}(\lambda _i\lambda ),$$ (6) $`\lambda _2,\lambda _3,\mathrm{},\lambda _{|V|}`$ being the remaining eigenvalues. Now the sum of all elements in a row of the determinant $`\left|M_{ij}(G)\lambda \delta _{ij}\right|`$ is $`\lambda `$. This permits us to replace the first column of $`\mathrm{det}\left|M_{ij}(G)\lambda \delta _{ij}\right|`$ by a column of elements $`\lambda `$ without affecting its value. Next we carry out a Laplace expansion of the resulting determinant along the modified column, obtaining $$\mathrm{det}\left|M_{ij}(G)\lambda \delta _{ij}\right|=\lambda \underset{i=1}{\overset{|V|}{}}C_{i1}(\lambda ),$$ (7) where $`C_{i1}(\lambda )`$ is the cofactor of the ($`i1`$)-th element of the determinant. Combining Eqs. (5)-(7), we are led to the identity $$F(\lambda )=\underset{i=1}{\overset{|V|}{}}C_{i1}(\lambda ).$$ (8) Now, $`C_{i1}(0)`$ is precisely the cofactor of the ($`i1`$)-th element of $`𝐌(G)`$ which, by Eq. (4), is equal to the tree generating function $`T(G;\{x_{ij}\})`$. It follows that, after setting $`\lambda =0`$ in Eq. (8), we obtain an expression giving the tree generating function in terms of the eigenvalues of the tree matrix \[7, p. 39\] $$T(G;\{x_{ij}\})=\frac{1}{|V|}\underset{i=2}{\overset{|V|}{}}\lambda _i.$$ (9) This result can also be deduced by considering the tree matrix of a graph obtained from $`G`$ by adding an auxiliary vertex connected to all vertices with edges of weight $`x`$, followed by taking the limit of $`x0`$ . ## 3 HYPERCUBIC LATTICES We now deduce the closed-form expression for the tree generating function for a hypercubic lattice in $`d`$ dimensions under various boundary conditions. 3.1. Free boundary conditions THEOREM 1. Let $`𝐙_d`$ be a $`d`$-dimensional hypercubic lattice of size $`N_1\times N_2\times \mathrm{}\times N_d`$ with edge weights $`x_i`$ along the $`i`$th direction, $`i=1,2,\mathrm{},d`$. The tree generating function for $`𝐙_d`$ is $`T(𝐙_d;\{x_i\})`$ $`=`$ $`{\displaystyle \frac{2^{𝒩1}}{𝒩}}{\displaystyle \underset{n_1=0}{\overset{N_11}{}}}\mathrm{}{\displaystyle \underset{n_d=0}{\overset{N_d1}{}}}\left[{\displaystyle \underset{i=1}{\overset{d}{}}}x_i\left(1\mathrm{cos}{\displaystyle \frac{n_i\pi }{N_i}}\right)\right],`$ $`(n_1,\mathrm{},n_d)(0,\mathrm{},0),`$ where $`𝒩=N_1N_2\mathrm{}N_d`$. PROOF. The tree matrix of $`𝐙_d`$ assumes the form of a linear combination of direct products of smaller matrices, $`𝐌(𝐙_d)`$ $`=`$ $`{\displaystyle \underset{i=1}{\overset{d}{}}}x_i[2I_{N_1}I_{N_2}\mathrm{}I_{N_d}`$ (11) $``$ $`I_{N_1}\mathrm{}I_{N_{i1}}H_{N_i}I_{N_{i+1}}\mathrm{}I_{N_d}],`$ where $`I_N`$ is an $`N\times N`$ identity matrix and $`H_N`$ is the $`N\times N`$ tri-diagonal matrix $$H_N=\left(\begin{array}{cccccccc}1& 1& 0& 0& \mathrm{}& 0& 0& 0\\ 1& 0& 1& 0& \mathrm{}& 0& 0& 0\\ 0& 1& 0& 1& \mathrm{}& 0& 0& 0\\ \mathrm{}& \mathrm{}& \mathrm{}& \mathrm{}& \mathrm{}& \mathrm{}& \mathrm{}& \mathrm{}\\ 0& 0& 0& 0& \mathrm{}& 1& 0& 1\\ 0& 0& 0& 0& \mathrm{}& 0& 1& 1\end{array}\right).$$ (12) It is readily verified that $`H_N`$ is diagonalized by the similarity transformation $$S_NH_NS_N^1=\mathrm{\Lambda }_N,$$ (13) where $`S_N`$ and $`S_N^1`$ are $`N\times N`$ matrices with elements $`\left(S_N\right)_{mn}`$ $`=`$ $`\left(S_N^1\right)_{nm}=\sqrt{{\displaystyle \frac{2}{N}}}\mathrm{cos}\left[(2n+1)\left({\displaystyle \frac{m\pi }{2N}}\right)\right]+\left(\sqrt{{\displaystyle \frac{1}{N}}}\sqrt{{\displaystyle \frac{2}{N}}}\right)\delta _{m,0},`$ (14) $`m,n=0,1,\mathrm{},N1,`$ and $`\mathrm{\Lambda }_N`$ is an $`N\times N`$ diagonal matrix with diagonal elements $$\lambda _n=2\mathrm{cos}\frac{n\pi }{N},n=0,1,\mathrm{},N1.$$ (15) Here $`\delta _{m,n}`$ is the Kronecker delta. It follows that $`𝐌(𝐙_d)`$ is diagonalized by the similarity transformation $$𝐒_𝒩𝐌(𝐙_d)𝐒_𝒩^1=\mathrm{\Lambda }_𝒩,$$ (16) where $$𝐒_𝒩=S_{N_1}S_{N_2}\mathrm{}S_{N_d},$$ (17) and $`\mathrm{\Lambda }_𝒩`$ is an $`𝒩\times 𝒩`$ diagonal matrix with diagonal elements $$\lambda _{n_1,\mathrm{},n_d}=2\underset{i=1}{\overset{d}{}}x_i\left[1\mathrm{cos}\frac{n_i\pi }{N_i}\right],n_i=0,1,\mathrm{},N_i1.$$ (18) Now, we have $`\lambda _{n_1,\mathrm{},n_d}=0`$ for $`n_1=n_2=\mathrm{}=n_d=0`$. This establishes Theorem 1 after using Eq. (9). Q.E.D. REMARK. The result Eq. (18) generalizes the $`d=2`$ eigenvalues of M$`(𝐙_2)`$ for $`x_i=1`$ reported in \[7, p. 74\]. 3.2. Periodic boundary conditions In applications in physics one often requires periodic boundary conditions depicted by the condition that two “boundary” vertices at coordinates $`(\mathrm{},n_i=1,\mathrm{})`$ and $`(\mathrm{},n_i=N_i,\mathrm{})`$, $`i=1,2,\mathrm{},d`$, are connected by an extra edge. This leads to a lattice $`𝐙_d^{\mathrm{Per}}`$ which is a regular graph with degree $`2d`$ at all vertices. For $`d=2`$, for example, $`𝐙_2^{\mathrm{Per}}`$ can be regarded as being embedded on the surface of a torus. THEOREM 2. Let $`𝐙_d^{\mathrm{Per}}`$ be a hypercubic lattice in $`d`$ dimensions of size $`N_1\times N_2\times \mathrm{}\times N_d`$ with edge weights $`x_i`$ along the $`i`$th direction, $`i=1,2,\mathrm{},d`$ with periodic boundary conditions. The tree generating function for $`𝐙_d^{\mathrm{Per}}`$ is $`T(𝐙_d^{\mathrm{Per}};\{x_i\})`$ $`=`$ $`{\displaystyle \frac{2^{𝒩1}}{𝒩}}{\displaystyle \underset{n_1=0}{\overset{N_11}{}}}\mathrm{}{\displaystyle \underset{n_d=0}{\overset{N_d1}{}}}\left[{\displaystyle \underset{i=1}{\overset{d}{}}}x_i\left(1\mathrm{cos}{\displaystyle \frac{2n_i\pi }{N_i}}\right)\right],`$ $`(n_1,\mathrm{},n_d)(0,\mathrm{},0).`$ PROOF. The tree matrix assumes the form $`𝐌(𝐙_d^{\mathrm{Per}})`$ $`=`$ $`{\displaystyle \underset{i=1}{\overset{d}{}}}x_i[2I_{N_1}I_{N_2}\mathrm{}I_{N_d}`$ (20) $``$ $`I_{N_1}\mathrm{}I_{N_{i1}}G_{N_i}I_{N_{i+1}}\mathrm{}I_{N_d}],`$ where $`G_N`$ is the $`N\times N`$ cyclic matrix $$G_N=\left(\begin{array}{cccccccc}0& 1& 0& 0& \mathrm{}& 0& 0& 1\\ 1& 0& 1& 0& \mathrm{}& 0& 0& 0\\ 0& 1& 0& 1& \mathrm{}& 0& 0& 0\\ \mathrm{}& \mathrm{}& \mathrm{}& \mathrm{}& \mathrm{}& \mathrm{}& \mathrm{}& \mathrm{}\\ 0& 0& 0& 0& \mathrm{}& 1& 0& 1\\ 1& 0& 0& 0& \mathrm{}& 0& 1& 0\end{array}\right).$$ (21) As in Eq. (16), the matrix $`𝐌(𝐙_d^{\mathrm{Per}})`$ can be diagonalized by a similarity transformation generated by $$𝐑_𝒩=R_{N_1}R_{N_2}\mathrm{}R_{N_d},$$ (22) where $`R_N`$ is an $`N\times N`$ matrix with elements $$\left(R_N\right)_{nj}=\left(R_N^1\right)_{jn}^{}=N^{1/2}e^{i2\pi jn/N},$$ (23) where denotes the complex conjugate, yielding eigenvalues of $`G_N`$ as $$\lambda _n=2\mathrm{cos}\frac{2n\pi }{N},n=0,1,\mathrm{},N1.$$ (24) This establishes Theorem 2 after using Eq. (9). Q.E.D. 3.3. Periodic boundary conditions along $`md`$ directions COROLLARY. Let $`𝐙_d^{\mathrm{Per}(m)}`$ be a hypercubic lattice in $`d`$ dimensions of size $`N_1\times N_2\times \mathrm{}\times N_d`$ with periodic boundary conditions in directions $`1,2,\mathrm{},md`$ and free boundaries in the remaining $`dm`$ directions. The tree generating function is $`T(𝐙_d^{\mathrm{Per}(m)};\{x_i\})={\displaystyle \frac{2^{𝒩1}}{𝒩}}{\displaystyle \underset{n_1=0}{\overset{N_11}{}}}\mathrm{}{\displaystyle \underset{n_d=0}{\overset{N_d1}{}}}[{\displaystyle \underset{i=1}{\overset{m}{}}}x_i(1\mathrm{cos}{\displaystyle \frac{2n_i\pi }{N_i}})`$ $`+{\displaystyle \underset{i=m+1}{\overset{d}{}}}x_i(1\mathrm{cos}{\displaystyle \frac{n_i\pi }{N_i}})],(n_1,\mathrm{},n_d)(0,\mathrm{},0).`$ (25) ## 4 THE MÖBIUS STRIP AND THE KLEIN BOTTLE Due to the interplay with the conformal field theory , it is of current interest in statistical physics to study lattice systems on non-orientable surfaces . Here, we consider two such surfaces, the Möbius strip and the Klein bottle, and obtain the respective tree generating functions. 4.1. The Möbius strip THEOREM 3. Let $`𝐙_2^{\mathrm{Mob}}`$ be an $`M\times N`$ simple quartic net embedded on a Möbius strip forming a Möbius net of width $`M`$ and twisted in the direction $`N`$, with edge weights $`x_1`$ and $`x_2`$ along directions $`M`$ and $`N`$ respectively. The tree generating function for $`𝐙_2^{\mathrm{Mob}}`$ is $`T(𝐙_2^{\mathrm{Mob}};\{x_1,x_2\})={\displaystyle \frac{2^{MN1}}{MN}}{\displaystyle \underset{m=0}{\overset{M1}{}}}{\displaystyle \underset{n=0}{\overset{N1}{}}}[x_1(1\mathrm{cos}{\displaystyle \frac{m\pi }{M}})`$ $`x_2(1\mathrm{cos}{\displaystyle \frac{4n3(1)^m}{2N}}\pi )],(m,n)(0,0).`$ (26) PROOF. Specifically, let the the two vertices at coordinates $`\{m,1\}`$ and $`\{Mm,N\},m=1,2,\mathrm{},M`$ be connected with a lattice edge of weight $`x_2`$. Then the tree matrix assumes the form $$𝐌(𝐙_2^{\mathrm{Mob}})=2(x_1+x_2)I_MI_Nx_1H_MI_Nx_2\left[I_MF_N+J_MK_N\right],$$ (27) where $$F_N=\left(\begin{array}{cccccccc}0& 1& 0& 0& \mathrm{}& 0& 0& 0\\ 1& 0& 1& 0& \mathrm{}& 0& 0& 0\\ 0& 1& 0& 1& \mathrm{}& 0& 0& 0\\ \mathrm{}& \mathrm{}& \mathrm{}& \mathrm{}& \mathrm{}& \mathrm{}& \mathrm{}& \mathrm{}\\ 0& 0& 0& 0& \mathrm{}& 1& 0& 1\\ 0& 0& 0& 0& \mathrm{}& 0& 1& 0\end{array}\right),J_M=\left(\begin{array}{cccccc}0& 0& \mathrm{}& 0& 0& 1\\ 0& 0& \mathrm{}& 0& 1& 0\\ 0& 0& \mathrm{}& 1& 0& 0\\ \mathrm{}& \mathrm{}& \mathrm{}& \mathrm{}& \mathrm{}& \mathrm{}\\ 0& 1& \mathrm{}& 0& 0& 0\\ 1& 0& \mathrm{}& 0& 0& 0\end{array}\right),$$ $$K_N=\left(\begin{array}{cccccc}0& 0& 0& \mathrm{}& 0& 1\\ 0& 0& 0& \mathrm{}& 0& 0\\ 0& 0& 0& \mathrm{}& 0& 0\\ \mathrm{}& \mathrm{}& \mathrm{}& \mathrm{}& \mathrm{}& \mathrm{}\\ 0& 0& 0& \mathrm{}& 0& 0\\ 1& 0& 0& \mathrm{}& 0& 0\end{array}\right).$$ Since $`H_M`$ and $`J_M`$ commute, they can be simultaneously diagonalized by applying the similarity transformation Eq. (13). The transformed matrix $`𝐒_𝒩𝐌(𝐙_2^{\mathrm{Mob}})𝐒_𝒩^1`$ is “block diagonal” with $`N\times N`$ blocks $$2\left(x_1x_1\mathrm{cos}\frac{m\pi }{M}+x_2\right)I_Nx_2\left(F_N+(1)^mK_N\right),m=0,1,\mathrm{},M1.$$ (28) Now, the eigenvalues of $`G_N=F_N+K_N`$ and $`F_NK_N`$ are, respectively, $`2\mathrm{cos}[2(n+1)\pi /N]`$ and $`2\mathrm{cos}[(2n+1)\pi /N],n=0,1,\mathrm{},N1`$. Theorem 3 is established by combining these results with Eq. (9). Q.E.D. REMARK. For $`M=2`$ and $`x_1=x_2=1`$, Eq. (26) gives the number of spanning trees on a $`2\times N`$ Möbius ladder as $`N_{SPT}`$ $`=`$ $`{\displaystyle \frac{1}{2N}}{\displaystyle \underset{j=1}{\overset{2N1}{}}}\left[3(1)^j2\mathrm{cos}{\displaystyle \frac{j\pi }{N}}\right]`$ (29) $`=`$ $`{\displaystyle \frac{N}{2}}\left[2+(2+\sqrt{3})^N+(2\sqrt{3})^N\right].`$ These two equivalent expressions have previously been given by \[7, p.218\] and by Guy and Harary , respectively. 4.2. The Klein bottle The embedding of an $`M\times N`$ simple quartic net on a Klein bottle is accomplished by further imposing a periodic boundary condition to $`𝐙_2^{\mathrm{Mob}}`$ in the M direction, namely, by connecting vertices of $`𝐙_2^{\mathrm{Mob}}`$ at coordinates $`\{1,n\}`$ and $`\{M,n\}`$, $`n=1,2,\mathrm{},N`$ with an edge of weight $`x_1`$. This leads to a lattice $`𝐙_2^{\mathrm{Klein}}`$ of the topology of a Klein bottle. THEOREM 4 The tree generating function for $`𝐙_2^{\mathrm{Klein}}`$ (described in the above) is $`T(𝐙_2^{\mathrm{Klein}};\{x_1,x_2\})={\displaystyle \frac{2^{MN1}}{MN}}\left[{\displaystyle \underset{n=1}{\overset{N1}{}}}2x_2\left(1\mathrm{cos}{\displaystyle \frac{2n\pi }{N}}\right)\right]`$ $`\times {\displaystyle \underset{m=1}{\overset{\left[\frac{M1}{2}\right]}{}}}{\displaystyle \underset{n=0}{\overset{2N1}{}}}[2x_1(1\mathrm{cos}{\displaystyle \frac{2m\pi }{M}})+2x_2(1\mathrm{cos}{\displaystyle \frac{n\pi }{N}})]`$ $`\times \{\begin{array}{cc}_{n=0}^{N1}\left[4z_12z_2\left(1\mathrm{cos}\frac{(2n+1)\pi }{N}\right)\right],\hfill & \text{for }M\text{ even}\hfill \\ 1,\hfill & \text{for }M\text{ odd,}\hfill \end{array}`$ (32) where $`[n]`$ is the integral part of $`n`$. PROOF. The tree matrix of $`𝐙_2^{\mathrm{Klein}}`$ assumes the form $$𝐌(𝐙_2^{\mathrm{Klein}})=2(x_1+x_2)I_MI_Nx_1G_MI_Nx_2\left[I_MF_N+J_MK_N\right].$$ (33) To obtain its eigenvalues, we first apply the similarity transformation generated by $`R_M`$ in the $`M`$ subspace. While this diagonalizes $`G_M`$ with eigenvalues $`2\mathrm{cos}(2m\pi /M)`$, $`m=0,1,\mathrm{},M1`$, it transforms the matrix $`J_M`$ into $$R_MJ_MR_M^1=\left(\begin{array}{ccccccc}1& 0& 0& \mathrm{}& 0& 0& 0\\ 0& 0& 0& \mathrm{}& 0& 0& \omega \\ 0& 0& 0& \mathrm{}& 0& \omega ^2& 0\\ \mathrm{}& \mathrm{}& \mathrm{}& \mathrm{}& \mathrm{}& \mathrm{}& \mathrm{}\\ 0& 0& \omega ^{M2}& \mathrm{}& 0& 0& 0\\ 0& \omega ^{M1}& 0& \mathrm{}& 0& 0& 0\end{array}\right),$$ (34) where $`\omega =e^{i2\pi /M}`$, and thus $`𝐌(𝐙_2^{\mathrm{Klein}})`$ into $$\left(\begin{array}{ccccccc}A_0+B_0& 0& 0& \mathrm{}& 0& 0& 0\\ 0& A_1& 0& \mathrm{}& 0& 0& B_1\\ 0& 0& A_2& \mathrm{}& 0& B_2& 0\\ \mathrm{}& \mathrm{}& \mathrm{}& \mathrm{}& \mathrm{}& \mathrm{}& \\ 0& 0& B_{M2}& \mathrm{}& 0& A_{M2}& 0\\ 0& B_{M1}& 0& \mathrm{}& 0& 0& A_{M1}\end{array}\right),$$ (35) where $`A_m`$ and $`B_m`$ are $`N\times N`$ matrices given by $`A_m`$ $`=`$ $`2\left[x_1+x_2x_1\mathrm{cos}{\displaystyle \frac{2m\pi }{M}}\right]I_Nx_2F_N,`$ $`B_m`$ $`=`$ $`\omega ^mx_2K_N,m=0,1,\mathrm{},M1.`$ (36) The matrix Eq. (35) is block diagonal with blocks $`U_N(0)`$ $`=`$ $`A_0+B_0=x_2(2I_NG_N)`$ $`U_{2N}(m)`$ $`=`$ $`\left(\begin{array}{cc}A_m& B_m\\ B_{Mm}& A_{Mm}\end{array}\right),m=1,2,\mathrm{},\left[{\displaystyle \frac{M1}{2}}\right]`$ (39) and for $`M=`$ even, $`U_N(M/2)`$ $`=`$ $`A_{M/2}+B_{M/2},`$ (40) $`=`$ $`2(2x_1+x_2)I_Nx_2(F_NK_N),`$ where the subscripts of the $`U`$ matrices denote the matrix dimensions. It follows that we need only to find the eigenvalues of the $`U`$ matrices. Eigenvalues of $`U_N(0)`$ and $`U_N(M/2)`$ can be deduced from those of $`G_N`$ and $`F_NK_N`$. Furthermore, eigenvalues of $`U_{2N}(m)`$ are obtained from those of $`G_N`$ after applying the similarity transformation $$T_{2N}(m)U_{2N}(m)T_{2N}^1(m)=2\left(x_1+x_2x_1\mathrm{cos}\frac{2m\pi }{m}\right)I_{2N}x_2G_{2N}$$ (41) where $$T_{2N}(m)=\left(\begin{array}{cc}I_N& 0\\ 0& \omega ^mI_N\end{array}\right).$$ (42) Combining these results with Eq. (9), we are led to Eq. (32) and the theorem. Q.E.D. ## Acknowledgment We thank T. K. Lee for the hospitality at the Center for Theoretical Sciences where this research is carried out. We are grateful to L. H. Kauffman for a useful conversation and to R. Shrock for calling our attention to Refs. and . The work of FYW is supported in part by NSF Grant DMR-9614170, and the support of the National Science Council, Taiwan, is also gratefully acknowledged.
warning/0001/astro-ph0001376.html
ar5iv
text
# 1 Introduction ## 1 Introduction The evolution of the space density of clusters of galaxies is a measurement sensitive to the physical and cosmological parameters of models of structure formation. We describe a deep X-ray survey of clusters of galaxies (the Wide Angle ROSAT Pointed Survey or WARPS), and use it to measure the evolution of the X-ray luminosity function (XLF) of clusters of galaxies , , ). Fairley et al (these proceedings) describe our results on the evolution of the cluster X-ray luminosity-temperature relation at high redshifts . ## 2 Survey method Our goal was to compile a complete, unbiased, X-ray selected sample of clusters of galaxies from serendipitous detections in deep, high-latitude ROSAT PSPC pointings. A detailed review of the X-ray source detection algorithm used (Voronoi Tessellation and Percolation or VTP), the sample selection and flux correction techniques are given in . VTP is particularly well suited to the detection and characterization of low surface brightness emission and to the recognition of extended sources. Only the centre of the ROSAT PSPC field was used, up to an off-axis angle of 15 arcmin. However, even at this relatively small off-axis angle, the PSPC point-spread function degrades to $``$55 arcsec (FWHM, at 1 keV) and there is a possibility that some clusters at the edge of the PSPC fields (where most of the survey area is) would remain unresolved. To reduce this possible incompleteness, our optical follow-up observations are not limited to extended X-ray sources but also include likely point X-ray sources without obvious optical counterparts. The flux limit is 5.5x10<sup>-14</sup> erg cm<sup>-2</sup> s<sup>-1</sup> (0.5-2 keV) and the total solid angle 73 deg<sup>2</sup>. Detailed simulations were performed to derive the survey sensitivity as a function of both source flux and source extent. ## 3 The X-ray luminosity function of clusters The survey currently contains over 150 clusters and groups of galaxies, approximately half of which are at redshifts z$`>`$0.3. Spectroscopic redshifts for all but a few have been obtained and the survey is complete at z$`<`$0.85. The remaining few distant cluster candidates are very probably all at higher redshifts, based on their brightest cluster galaxy magnitudes. The XLFs shown here are based on an initial analysis; while some details will change (eg correction for AGN contamination) in a final version, we do not expect there to be major changes. The cluster XLF is shown in Fig 1. A value of q<sub>0</sub>=0.5 has been assumed in calculating the luminosities and volumes but the results are qualitatively similar for q<sub>0</sub>=0.1. The WARPS data points extend over almost 3 decades of luminosity, from 10<sup>42</sup> erg s<sup>-1</sup> at low redshifts and up to 8x10<sup>44</sup> erg s<sup>-1</sup> (0.5-2 keV) at high redshifts. There is no evidence for evolution of the XLF at any redshift up to z=0.8 when compared with the local BCS XLF , shown as a solid line. The error bars are based on a Poissonian distribution of the number of clusters in each bin. The survey area for each cluster has been calculated using the cluster flux but making the simplifying assumption that all clusters have the same observed surface brightness profile (characterised as a constant angular core radius and $`\beta `$=2/3 index). Mean differences from the true survey areas are small ($`<`$10%). As the XLFs at z=0.3-0.5 and z=0.5-0.85 are consistent with each other, in the lower panel of the figure we have combined both redshift ranges to increase the statistical accuracy at the expense of a rather broad redshift bin. Nevertheless, the small error bars on the high redshift points (filled circles) indicate that any evolutionary factor $`>`$1.5 in the space density of moderate luminosity clusters (3x10$`{}_{}{}^{43}<`$L$`{}_{X}{}^{}<`$3x10<sup>44</sup> erg s<sup>-1</sup>) is ruled out. ## 4 Discussion Several recent deep X-ray clusters surveys are described in these proceedings and elsewhere. In the regime where these surveys have good statistical accuracy (ie. moderate X-ray luminosities $``$3x10<sup>43</sup>-3x10<sup>44</sup> erg s<sup>-1</sup>) there is a consensus that no evolution of the XLF is observed to z$``$0.7. Five surveys agree on this point (EMSS, RDCS, SHARC, CfA, WARPS) and the only disagreement (the RIXOS survey ) can be understood in terms of the RIXOS source detection algorithm. Most models of the growth of structure in the Universe with $`\mathrm{\Omega }_m`$=1 predict rapid evolution of the space density of clusters of galaxies (eg. scaling relation models or a cold+hot dark matter model ). The tight constraints on the level of evolution allowed by the observations rules out most of the $`\mathrm{\Omega }_m`$=1 models. In models with $`\mathrm{\Omega }_m`$0.3 (with or without a cosmological constant) the predicted evolution is much less rapid and these values of $`\mathrm{\Omega }_m`$ are consistent with the lack of evolution observed. At the higher X-ray luminosities of the most massive clusters (L$`{}_{X}{}^{}>`$5x10<sup>44</sup> erg s<sup>-1</sup>), there is some disagreement between the results of different surveys as to the degree of evolution found (if any), but this may partly be due to the small numbers of high luminosity clusters in any one survey. This is unfortunate, since it is the most massive clusters which have most leverage to constrain the cosmological parameters of model predictions. Wide area X-ray surveys, designed to detect the rare high luminosity, massive clusters in large numbers at high redshifts (z$`>`$0.5) are planned (see Lumb & Jones and Ebeling et al in these proceedings). Future work will concentrate on the handful of clusters in the WARPS survey at z$``$1, on measurements of the evolution of the cluster X-ray temperature function using Chandra and XMM, and studies of the Butcher-Oemler effect in X-ray selected clusters. Acknowledgements. We acknowledge useful discussions with Pat Henry.
warning/0001/cond-mat0001079.html
ar5iv
text
# Bose condensates in a harmonic trap near the critical temperature ## I Introduction Since the first observations of Bose-Einstein condensation (BEC) in dilute alkali metal atom gases , experimental developments have posed many new tests for many-body theory, even though weakly interacting Bose gases have long been used as a textbook paradigm . Numerous theoretical approaches have been employed in order to obtain accurate results for both the ground-state and non-equilibrium properties of the trapped Bose systems . However, there have been notable differences between theoretical results and experimental data on the excitation frequencies near the transition temperature $`T_c`$ . This problem has inspired the introduction of a renormalized effective atom-atom interaction . Recently developed theoretical approaches that incorporate the dynamics of the noncondensate density but without a renormalized interaction have resulted in excitation frequencies in closer agreement with experiment. Nevertheless, the unresolved issues for Bose systems near $`T_c`$ has provided a motivation for us to examine further the theoretical and numerical methods for modeling confined Bose gases near $`T_c`$. We have numerically implemented the most plausible and tractable equilibrium mean-field theories in order to systematically survey various properties of these systems. In this work, we follow the standard mean-field theory , with certain modifications described in detail below. The nonlinear Gross-Pitaevskii (GP) equation, which includes interactions between the condensate and the thermal atoms, is solved for a static condensate containing $`N_c`$ atoms. The eigenvalue of the GP equation, $`\stackrel{~}{\mu }`$, is usually identified with the thermodynamic chemical potential $`\mu `$. The linear response of the system is represented by the Hartree-Fock-Bogoliubov (HFB) equations, which yield the quasiparticle energies and amplitudes. These in turn determine the number of noncondensed atoms $`N_T`$ as well as various coherence terms (thermodynamic averages over two or more Bose field operators). The GP and HFB equations are iterated to self-consistency at a given temperature $`T`$, subject to a fixed total number of atoms in the system $`N=N_c+N_T`$. As emphasized by Griffin , the coherence terms yield an excitation spectrum that is not gapless: the lowest-energy mode of the HFB equations has finite energy and does not coincide with the solution of the GP equation. The HFB-Popov (HFBP) approximation, which neglects these terms, has been quite successful in describing the properties of the trapped Bose gases, but is not well-grounded theoretically, and fails to yield accurate predictions for the low-lying excitations at high temperatures . In this work, we explore a recently proposed extension of the HFBP theory that incorporates the coherence terms in a gapless manner ; in addition, we modify the commonly used identification of the chemical potential with the eigenvalue of the GP equation. The identification of the chemical potential with the eigenvalue of the GP equation is incorrect in general. In the grand canonical ensemble, the chemical potential is defined as $`\mu =E/N`$, corresponding to the energy cost $`E`$ of adding a particle to the entire system, not only to the condensate. For a dilute, weakly interacting Bose gas at $`T=0`$, for which the population of noncondensed states (the depletion) is negligible, the identification $`\stackrel{~}{\mu }=\mu `$ is justified. At finite temperatures, however, the assumption yields results that are discontinuous as a function of the $`s`$-wave scattering length $`a_{\mathrm{sc}}`$. To a better approximation, we find that the chemical potential at finite temperatures is given by the eigenvalue of the GP equation plus a term that varies inversely with the number of condensate atoms. The resulting equations provide an improved description of these finite systems, yielding observables that are both continuous with $`a_{\mathrm{sc}}`$ and similar to those obtained using path integral Monte Carlo techniques . It is presently unclear to what extent many-body effects beyond the mean-field approximation modify the effective interactions among Bose-condensed atoms in harmonic traps . For the homogeneous Bose gas, it is now established both from renormalization group and perturbation theories that the many-body T-matrix, or effective s-wave scattering length $`a`$, goes to zero at $`T_c`$. The low-energy, long wavelength limit of the many-body T-matrix has been shown to be closely related to the coherence term $`m_T`$ ; this ‘anomalous average’ represents two-particle correlations and is the Bose analogue of the superconducting order parameter in interacting Fermi systems. Renormalizing the interaction using the $`m_T`$ yields a gapless HFB theory without having to invoke the Popov approximation , but it remains uncertain whether the prescription is appropriate for nonuniform systems. The implications of this theory for trapped Bose condensates are explored numerically below, and the results are compared to those obtained with the Popov approximation and path-integral Monte Carlo methods. In view of these somewhat conflicting results and unresolved issues, there is strong motivatation for continued development of numerical methods in order to implement various models and obtain quantitative predictions for comparison with experiment. Quantum Monte Carlo methods are able to provide accurate results for certain observable quantities. The computational procedure is lengthy, however, and is not demonstrably able to yield excitation frequencies since it typically applies only to equilibrium configurations. Local density approximations (LDA) are much simpler to apply, but the standard forms fail near $`T_c`$ and are questionable when the density is so small that the local collision rate is insufficient to establish local thermodynamic equilibrium. On the other hand, widely used basis set techniques are generally unable to represent the large numbers of atoms in excited states at high-temperatures. Recently, Reidl et al. have used (for 2,000 Rb atoms at $`T=0.5T_c`$) a hybrid method in which a sum over discrete quasiparticle states at low energies is supplemented by an integral over an energy-dependent LDA above some cutoff energy. The interactions of these two subensembles with each other are expressed by mean-field potentials that represent the effect of background atoms. In the present work, the low-lying states are obtained by solving the HFB equations using the discrete variable representation (DVR) and the cutoff energy is raised until the results converge to within a stated tolerance. The techniques employed have enabled the investigation of trapped Bose gases at finite-temperatures containing a larger number of atoms than in previous calculations that we are aware of. As a result, the approach of these systems to the local thermodynamic equilibrium and to the hydrodynamic limit can be explored. In Section II A, we outline the GP and HFB equations. We discuss the chemical potential and gapless theories in Sections II B and II C, respectively. Section II D reviews LDA methods both as alternative approaches for comparison purposes, and the complementary use for the most energetic atoms. In Section III, we discuss our numerical methods and iteration procedures. Section IV presents results for Bose atoms in a spherically symmetric harmonic trap as a function of the scaled $`s`$-wave scattering length, total number of atoms, and temperature. ## II Theoretical framework ### A Thermal sums over quasiparticle states The derivation of mean-field equations for a weakly interacting, dilute Bose gas has been described in detail elsewhere . The question of the chemical potential for $`T>0`$ for thermal sums of quasiparticle states deserves more thorough discussion, however, and we modify the standard procedure. In addition, following discussions by the Burnett et al. , we treat the anomalous (coherence) terms $`m_T`$ in a manner that produces a ‘gapless’ theory. Following the standard approach, we decompose the Bose field operator into a $`c`$-number for the condensate, plus an operator representing its fluctuations. The full many-body Hamiltonian is approximated using mean-field theory, becoming explicitly number-nonconserving. The grand canonical ensemble is used, and thus the chemical potential, $`\mu `$, and temperature, $`T`$, are the sole fixed quantities. The generalized Gross-Pitaevskii (GP) equation for the condensate and coupled Bogoliubov equations for the excited quasiparticle states are then solved. For a finite number atoms in a harmonic potential, however, the standard approach yields values for the mean condensate number $`N_c`$ that are discontinuous as a function of interaction strength $`a_{\mathrm{sc}}`$. In our approach, the eigenvalue of the GP equation, $`\stackrel{~}{\mu }`$, is determined by the mean number of atoms in the condensate $`N_c`$. In contrast, the chemical potential, $`\mu `$, is adjusted so that the mean total number of atoms is the desired value. A simple relationship is found connecting $`\stackrel{~}{\mu }`$, $`\mu `$, and $`N_c`$, which is adapted from the ideal Bose gas case. The Hamiltonian for an interacting Bose gas in a trap in the grand canonical ensemble is $`\widehat{H}\mu \widehat{N}`$ $`=`$ $`{\displaystyle }d𝐫\{\widehat{\psi }^{}[{\displaystyle \frac{\mathrm{}^2}{2M}}^2+V_{\text{ext}}\mu ]\widehat{\psi }`$ (2) $`+{\displaystyle \frac{g}{2}}\widehat{\psi }^{}\widehat{\psi }^{}\widehat{\psi }\widehat{\psi }\},`$ where the field operator $`\widehat{\psi }(𝐫)`$ satisfies $`[\widehat{\psi }(𝐫_1),\widehat{\psi }^{}(𝐫_2)]=\delta (𝐫_1𝐫_2)`$. The pseudopotential atom-atom interaction has been chosen to be $`V(𝐫_1𝐫_2)=g\delta (𝐫_1𝐫_2)`$, where the coupling constant $`g=4\pi \mathrm{}^2a_{sc}/M`$ is written in terms of the scattering length $`a_{\mathrm{sc}}`$ and mass $`M`$. The harmonic potential is $`V_{\text{ext}}=\frac{1}{/}2M\omega _0^2r^2`$ with trapping frequency $`\omega _0`$ is assumed to be isotropic. The Hamiltonian may be rewritten as $`\widehat{H}\mu \widehat{N}`$ $`=`$ $`H\stackrel{~}{\mu }\widehat{N}+\left(\stackrel{~}{\mu }\mu \right)\widehat{N}`$ (3) $`=`$ $`\widehat{K}+\left(\stackrel{~}{\mu }\mu \right)\widehat{N},`$ (4) where, as mentioned above, the Lagrange multiplier $`\stackrel{~}{\mu }`$ is related to the number of atoms in the condensate. In the following, we choose to diagonalize the operator $`\widehat{K}=\widehat{H}\stackrel{~}{\mu }\widehat{N}`$ rather than the original $`\widehat{H}\mu \widehat{N}`$; both choices must lead to the same excitation spectrum, though with a temperature-dependent shift of the vacuum for quasiparticle excitations. In order to make further progress, the Bose field operator $`\widehat{\psi }=\mathrm{\Phi }+\widehat{\varphi }`$ is now decomposed into $`\mathrm{\Phi }`$, a $`c`$-number for the condensate, and $`\widehat{\varphi }(𝐫)`$, which annihilates a thermal atom at $`𝐫`$. The condensate density is defined by $`n_c=|\mathrm{\Phi }|^2`$, and the number of condensate atoms is $`N_c=𝑑𝐫|\mathrm{\Phi }(𝐫)|^2`$. The noncondensate (thermal) density $`n_T`$ and anomalous (coherence) terms $`m_T`$ and $`\stackrel{~}{m}_T`$ are $$n_T=\widehat{\varphi }^{}\widehat{\varphi };m_T=\widehat{\varphi }\widehat{\varphi };\stackrel{~}{m}_T=\widehat{\varphi }^{}\widehat{\varphi }^{},$$ (5) where the brackets indicate a thermal average in the grand canonical ensemble, discussed in more detail below. The mean field approximation is used to reduce the third and fourth order terms to, respectively, first and second order in $`\widehat{\varphi },\widehat{\varphi }^{}`$ so that the Hamiltonian $`\widehat{K}`$ can be diagonalized, following the procedure normally used for $`\widehat{H}\mu \widehat{N}`$ . Excluding the possibility of aggregate motion and vortices , $`\mathrm{\Phi }`$ may be taken to be real. The first order terms (plus third order terms in mean-field approximation) in $`\widehat{K}`$ vanish if the equation for the condensate is taken to be the generalized GP equation: $$\left[\frac{\mathrm{}^2}{2M}^2+V_{\text{ext}}+g[n_c+2n_T+\stackrel{~}{m}_T]\right]\mathrm{\Phi }=\stackrel{~}{\mu }\mathrm{\Phi }.$$ (6) Note that $`\stackrel{~}{\mu }`$ is the eigenvalue of the GP equation. The part of $`\widehat{K}`$ that is zeroth order in the excited orbitals is a $`c`$-number $`K_0`$ $`=`$ $`{\displaystyle }d𝐫\mathrm{\Phi }(𝐫)[{\displaystyle \frac{\mathrm{}^2}{2M}}^2+V_{ext}\stackrel{~}{\mu }`$ (8) $`+{\displaystyle \frac{g}{2}}|\mathrm{\Phi }(𝐫)|^2]\mathrm{\Phi }(𝐫).`$ The terms in $`\widehat{K}`$ that are second order in $`\widehat{\varphi }`$ are (in the mean-field approximation) diagonalized by the canonical transformation $`\widehat{\varphi }(𝐫)={\displaystyle \underset{j}{}}[u_j(𝐫)\widehat{\alpha }_j+v_j^{}(𝐫)\widehat{\alpha }_j^{}]`$ (9) $`\widehat{\varphi }^{}(𝐫)={\displaystyle \underset{j}{}}[u_j^{}(𝐫)\widehat{\alpha }_j^{}+v_j(𝐫)\widehat{\alpha }_j],`$ (10) such that $`[\widehat{\alpha }_i,\widehat{\alpha }_j^{}]=\delta _{i,j}`$. The operator $`\widehat{K}`$ is diagonal to second order in $`\widehat{\varphi }`$ if the quasiparticle amplitudes $`u_j(𝐫)`$ and $`v_j(𝐫)`$ are solutions of the Bogoliubov coupled equations $`\widehat{}u_j(𝐫)+𝒬(𝐫)v_j(𝐫)`$ $`=`$ $`ϵ_ju_j(𝐫)`$ (11) $`\widehat{}v_j(𝐫)+𝒬(𝐫)u_j(𝐫)`$ $`=`$ $`ϵ_jv_j(𝐫),`$ (12) where $`\widehat{}=K+V_{\text{ext}}\stackrel{~}{\mu }+2gn(𝐫)`$, $`𝒬=g[n_c(𝐫)+m_T(𝐫)]`$, the total density is $`n(𝐫)=n_c(𝐫)+n_T(𝐫)`$. For ‘gapless’ theories, discussed further below, the $`j=0`$ ‘Goldstone mode,’ has the property $`ϵ_0=0`$, so that $`u_0(𝐫)=v_0(𝐫)=\mathrm{\Phi }(𝐫)`$. Thus on the $`ϵ_j`$ energy scale, the condensate has zero energy, and defines the vacuum for quasiparticle excitations. After the substitution, $`\widehat{\psi }=\mathrm{\Phi }+\widehat{\varphi }`$, the number operator $`\widehat{N}=𝑑𝐫\widehat{\psi }^{}(𝐫)\widehat{\psi }(𝐫)`$ contains terms, such as $`𝑑𝐫\mathrm{\Phi }\widehat{\varphi }`$, that do not conserve particle number. The Bogoliubov transformation (10) and coupled equations (12) introduce a quasiparticle basis such that terms $`\widehat{\alpha }_j^{}\widehat{\alpha }_j^{}`$ and $`\widehat{\alpha }_j\widehat{\alpha }_j`$ are eliminated, so quasiparticle number is conserved . The diagonalized Hamiltonian explicitly does not conserve particle number, however; the operator $`\widehat{K}`$ in the quasiparticle basis does not commute with the excited particle number operator $`\widehat{\varphi }^{}\widehat{\varphi }`$, which has contributions from $`\widehat{\alpha }^{}\widehat{\alpha }^{}`$ and $`\widehat{\alpha }\widehat{\alpha }`$ terms. In the grand canonical ensemble, only $`T`$ and $`\mu `$ are precisely defined, and all observables must be defined in terms of thermal averages. Each occupation number, including the condensate number, fluctuates about its mean value $$\widehat{N}_j\widehat{\alpha }^{}\widehat{\alpha },j=0,1,\mathrm{},$$ (13) where the explicit definition of the average $`\widehat{O}`$ is yet undefined. Similarly, both the eigenvalue $`\stackrel{~}{\mu }`$ of the GP equation (6) and the total energy $`E`$ fluctuate about their mean values. Inserting the transformation (10) into Eqs. (5) and introducing the identification given by Eq. (13), the normal and anomalous densities become $$n_T(𝐫)=\underset{j=1}{}\left\{\widehat{N}_j[|u_j(𝐫)|^2+|v_j(𝐫)|^2]+|v_j(𝐫)|^2\right\};$$ (14) $$m_T(𝐫)=\underset{j=1}{}u_j(𝐫)v_j^{}(𝐫)[2\widehat{N}_j+1].$$ (15) The standard normalization $`𝑑𝐫[|u_j(𝐫)|^2|v_j(𝐫)|^2]=1`$, yields $$𝑑𝐫[|u_j(𝐫)|^2+|v_j(𝐫)|^2]1+2V_j,$$ (16) where $`V_j=𝑑𝐫|v_j(𝐫)|^2`$. The quantities $`V_j`$ are related to the $`T=0`$ depletion, which is $`_{j=1}V_j`$. The relation between the total atom number and the quasiparticle occupation numbers is therefore $`\widehat{N}`$ $``$ $`N_c+N_T=N_0+{\displaystyle 𝑑𝐫n_T(𝐫)}`$ (17) $`=`$ $`N_c+{\displaystyle \underset{j=1}{}}[\widehat{N}_j(1+2V_j)+V_j],`$ (18) where the average number of atoms is written in terms of a contribution from the condensate and noncondensate (excitations). The thermal average of the diagonalized Hamiltonian then becomes $`\widehat{H}`$ $``$ $`\mu \widehat{N}=K_0+{\displaystyle }_{j=1}ϵ_j(\widehat{N}_jV_j)+(\stackrel{~}{\mu }\mu )\widehat{N}`$ (19) $`=`$ $`E_c\mu N_c+{\displaystyle \underset{j=1}{}}\{\widehat{N}_j[ϵ_j+(\stackrel{~}{\mu }\mu )(1+2V_j)]`$ (21) $`+V_j(\stackrel{~}{\mu }\mu ϵ_j)\},`$ where $`E_c=K_0+\stackrel{~}{\mu }N_c`$ is the total ground state, or condensate, energy. ### B Occupation factors and the chemical potential In the Bogoliubov approach, the ensemble is considered to be the sum of a condensate plus non-interacting quasiparticles. The mean occupation numbers of the quasiparticle states are to be determined from the grand partition function, $$\mathrm{\Omega }=\mathrm{Tr}\{\mathrm{exp}[\beta (\widehat{H}\mu \widehat{N})]\},$$ (22) through the standard identities $$N=\frac{1}{\beta }\left(\frac{\mathrm{ln}\mathrm{\Omega }}{\mu }\right)_T;E=\left(\frac{\mathrm{ln}\mathrm{\Omega }}{\beta }\right)_{\mu ,T}.$$ (23) Unfortunately, while the diagonalized Hamiltonian is written in terms of non-interacting single-quasiparticle energies, the expressions (18) and (21) involve the thermal averages of particle occupation that we are now seeking to determine. Furthermore, the factorization that one makes for an ideal Bose gas is invalid for a gas of interacting Bose atoms because the quasiparticle energies depend on the occupation numbers, as well as the reverse. Thus, rigorously, these occupation factors should be calculated self-consistently, along with Eqs. (6) and (12), since they depend on as well as determine the quasiparticle eigenvalues . To do so analytically would be a truly daunting task. We make several simplying assumptions in order to obtain results, but we emphasize that these questions merit further study. In reality, the probabilities $`\widehat{N}`$ will be peaked at the most probable values, as discussed below for the condensate. Therefore, when evaluating the sum over $`N_j`$ in Eq. (22), deviations of $`N_j^{}`$ from $`\widehat{N}_j^{}`$ for $`j^{}j`$ will not greatly modify the spectrum of the quasiparticle states. If this is so, a reasonable approximation is to replace $`\widehat{N}_j`$ by $`N_j`$ when estimating the mean occupation numbers from the grand partition function. If the dependence of $`ϵ_j`$ and $`N_j`$ on $`N_j^{}(jj^{})`$ is also neglected, then $`\mathrm{\Omega }`$ can be factored, and we obtain $`N_j`$ $`=`$ $`{\displaystyle \frac{\underset{N_j}{}N_j\mathrm{exp}\left\{\beta \left[ϵ_j+(\stackrel{~}{\mu }\mu )(1+2V_j)\right]N_j\right\}}{_{N_j}\mathrm{exp}\left\{\beta \left[ϵ_j+(\stackrel{~}{\mu }\mu )(1+2V_j)\right]N_j\right\}}}`$ (24) $``$ $`{\displaystyle \frac{1}{\mathrm{exp}\left\{\beta \left[ϵ_j+(\stackrel{~}{\mu }\mu )(1+2V_j)\right]\right\}1}},j.`$ (25) In order to obtain the result on the second line of Eq. (25), the population-dependences of the GP eigenvalue $`\stackrel{~}{\mu }`$ and the quasiparticle energies $`ϵ_j`$ are ignored. At sufficiently low temperatures, the $`ϵ_j`$ for trapped Bose-condensates are relatively insensitive to the value of $`N_c`$ and the temperature; indeed, in the Thomas-Fermi (TF) limit, valid for large condensates, the excitation frequencies at zero temperature are independent of $`N_c`$. Neglecting the factors $`V_j`$, and shifting the energy scale so that $`E_jϵ_j+\stackrel{~}{\mu }`$, one recovers the more conventional expression $$N_j=\frac{1}{\mathrm{exp}[\beta (E_j\mu )]1}.$$ (26) From this expression (26) for $`j=0`$, with $`E_0=\stackrel{~}{\mu }`$, one finds that the chemical potential $`\mu `$ and the eigenvalue of the GP equation $`\stackrel{~}{\mu }`$ are related by the expression $$\mu =\stackrel{~}{\mu }\frac{1}{\beta }\mathrm{ln}\left(1+\frac{1}{N_c}\right),N_c>0.$$ (27) For $`T=0`$ this gives the usual definition $`\stackrel{~}{\mu }=\mu `$, but for $`T>0`$ there is a correction to $`\mu `$ that increases as $`N_c`$ decreases. While this additional term will not be correct at high temperatures where the condensate is strongly depleted, it will be shown below that results obtained with this procedure are continuous functions of $`a_{\mathrm{sc}}`$ at all temperatures, while with $`\mu =\stackrel{~}{\mu }`$ they are not. It is difficult to go beyond the above approximations, but we will suggest possible avenues to proceed in future work. The major effect omitted is the dependence of the quasiparticle energies, $`E_j`$ (including $`E_0=\stackrel{~}{\mu }`$) on $`N_c`$. One can first consider the condensate term itself. We assume, for the moment, that factorization of $`\mathrm{\Omega }`$ (22) is valid, and write $$\mathrm{\Omega }=\mathrm{\Omega }_c\mathrm{\Omega }_T;\mathrm{\Omega }_c=\underset{N_c}{}e^{\beta (\mu N_cE_c)}$$ (28) In the Thomas-Fermi approximation (kinetic energy in the GP equation neglected), one obtains for a spherical condensate : $$\stackrel{~}{\mu }_{\mathrm{TF}}=\frac{1}{2}\left(\frac{15N_ca_{sc}}{a_0}\right)^{2/5}\gamma N_c^{2/5},$$ (29) where the harmonic oscillator length is $`a_0=\sqrt{\mathrm{}/M\omega }`$. The following relations follow in the same approximation: $$E_c=\frac{5}{7}\gamma N_c^{7/5}=\frac{5}{7}\stackrel{~}{\mu }_{\mathrm{TF}}N_c;\frac{E_c}{N_c}=\stackrel{~}{\mu }.$$ (30) Then from Eq. (28), neglecting $`\mathrm{\Omega }_T`$, one can obtain the mean value of the condensate occupation from $`N_c=_{N_c}N_cP(N_c)/_{N_c}P(N_c)`$, where $`P(N_c)=\mathrm{exp}[\beta (\mu N_c\frac{5}{/}7\gamma N_c^{7/5})]`$. We have verified numerically for typical values of $`\beta `$ and $`\mu `$ that the mean value is extremely close to the most probable value $`\overline{N}_c`$ for which $`P(N_c)`$ is maximum. Furthermore, an expansion of the exponent in the above expression for $`P(N_c)`$ yields a value for the variance of $`N_c`$, interpreted as the value of $`\sigma `$ in such that $`P(\overline{N}_c\pm \sigma )=(1/e)P(\overline{N}_c)`$. In the grand canonical ensemble at zero temperature, therefore, one obtains $`\delta N_c=\sqrt{(5/\beta \gamma )}\overline{N}_c^{3/10}`$, so that the fractional width of the occupation number distribution decreases as $`\overline{N}_c^{7/10}`$. This may be compared with the result of Giorgini et al., derived from excited state occupation numbers for the canonical ensemble, $`\delta N_c(T/T_c)N_c^{2/3}`$ . Either result confirms that the fluctuations in $`N_c`$ are relatively small for large $`N_c`$. One should next consider how the dependence of $`\mathrm{\Omega }_T`$ would effect $`N_c`$ and $`\delta N_c`$. This is left for future work. The dependence of the quasiparticle states on the occupation factors reflects the extensive nature of this finite interacting system; that is, adding a particle to the many-body system alters both the number and character of the accessible states. This behavior is similar to that of a finite gas of non-interacting particles obeying fractional exclusion statistics , which obey a statistics intermediate between that of bosons and fermions. The parameter representing the statistics has been identified with the strength of the delta-function potential for an interacting trapped Bose gas in two dimensions . Indeed, our expression (27) for the thermodynamic chemical potential is similar to that found for a non-interacting fractional-statistics gas at finite temperature . We hope to pursue these issues more fully in future work. ### C Gapless approximations We return to the conditions for ‘gaplessness.’ The GP (6) and Bogoliubov (12) equations together comprise the ‘Hartree-Fock-Bogoliubov’ (HFB) approximation for a dilute interacting Bose gas. In this case, one does not obtain $`ϵ_0=0`$, and the theory is said to be not gapless (the term has been taken from the homogeneous situation). In the Popov approximation, gaplessness is ensured by neglecting the coherence terms $`m_T`$ and $`\stackrel{~}{m}_T`$, but the justification for such an approximation is questionable . In order to convert HFB into a gapless theory and still retain the anomalous averages, Burnett et al. have recently proposed an alternative treatment in which the coupling functions for the condensate $`g_c(𝐫)`$ and excited states $`g_e(𝐫)`$ absorb the pairing correlations, and thereby take on a spatial dependence. Eq. (6) becomes $$\{K+V_{\text{ext}}+g_cn_c+2g_en_T\}\mathrm{\Phi }=\stackrel{~}{\mu }\mathrm{\Phi },$$ (31) and similarly, $`\widehat{}`$ and $`𝒬`$ appearing in Eqs. (12) become $$\widehat{}=K+V_{\mathrm{ext}}\stackrel{~}{\mu }+2g_cn_c+2g_en_T;𝒬=g_cn_c.$$ (32) In the proposed gapless theories, labeled G1, and G2, the coupling constants are chosen to be: $$\{g_c;g_e\}=\{\begin{array}{cc}\{g_1;g\},\hfill & \text{G1}\hfill \\ \{g_1;g_1\},\hfill & \text{G2}\hfill \end{array},$$ (33) where $$g_1(𝐫)=g\left[1+\frac{m_T(𝐫)}{n_c(𝐫)}\right].$$ (34) The renormalized coupling $`g_1`$ replaces the two-body T-matrix associated with binary atomic collisions, which is the scattering length $`a_{\mathrm{sc}}`$ in vacuo, by the zero momentum and energy limit of the homogeneous many-body T-matrix . In the G1 approximation, only the condensate-condensate and condensate-excited are dressed, while G2 is motivated by the expectation that all particle interactions should be similar. Renormalization of the coupling has the additional advantage of removing the ultraviolet divergence in $`m_T`$ resulting from high-energy quasiparticle contributions of Eq. (15) in the T-matrix approximation. In nonuniform systems such as the trapped Bose gases, however, the value of $`g_1(𝐫)`$ can diverge in regions near the condensate surface where the condensate density vanishes more rapidly than the anomalous average. In practice, this divergence may be eliminated by setting $`g_1(𝐫)=g\left[1+m_T(𝐫)/\left(n_c(𝐫)+\delta \right)\right]`$, where $`\delta 10^2`$. While the results, described in detail below, are found not to depend strongly on the choice of $`\delta `$, its existence underlines a deficiency in the theory in its present form. The consequences of the G1 approximation are not explored in this work. In the following, the notation $`g(𝐫)`$ will be used in place of $`g_1(𝐫)`$ and in distinction to $`g`$, which is unrenormalized. ### D Local Density Approximation In local density approximation (LDA) schemes, the condensate density is assumed to be varying sufficiently slowly that the population of excited states is determined entirely by the local potential and temperature. The thermal density may then be treated locally as if the interacting Bose gas were homogeneous. We will discuss three basic LDA schemes, and several variants. In the semiclassical approximation to the GP and HFB equations , the thermal atom quasiparticle amplitudes in the Bogoliubov equations (12) become local functions $`u(𝐩,𝐫)`$ and $`v(𝐩,𝐫)`$. With the Popov approximation, one obtains the coupled algebraic equations $$\left(\begin{array}{cc}(𝐩,𝐫)& gn_c(𝐫)\\ gn_c(𝐫)& (𝐩,𝐫)\end{array}\right)\left(\begin{array}{c}u(𝐩,𝐫)\\ v(𝐩,𝐫)\end{array}\right)=ϵ(𝐩,𝐫)\left(\begin{array}{c}u(𝐩,𝐫)\\ v(𝐩,𝐫)\end{array}\right),$$ (35) where $`(𝐩,𝐫)=p^2/2m+V_{\mathrm{ext}}(𝐫)\stackrel{~}{\mu }+2gn(𝐫)`$. With the condition $`u(𝐩,𝐫)^2v(𝐩,𝐫)^2=1`$, the local excitation energies may be immediately obtained, $`ϵ(𝐩,𝐫)=((𝐩,𝐫)^2g^2n_c^2(𝐫)^2)^{1/2}`$, and have the well known linear dispersion. The noncondensate density from Eq. (14) may then be easily found : $`n_T(𝐫)`$ $`=`$ $`{\displaystyle \frac{d^3𝐩}{(2\pi )^3}\left[\frac{(𝐩,𝐫)}{ϵ(𝐩,𝐫)}\left(n(𝐩,𝐫)+\frac{1}{2}\right)\frac{1}{2}\right]}`$ (37) $`\mathrm{\Theta }\left((𝐩,𝐫)^2g^2n_c^2(𝐫)\right),`$ where $$n(𝐩,𝐫)=\frac{1}{\mathrm{exp}[\beta (ϵ(𝐩,𝐫)+\stackrel{~}{\mu }\mu )]1},$$ (38) such that the theta function is unity when the argument is positive, and zero otherwise. These equations define the Hartree-Fock Bogoliubov Popov LDA, which we will refer to as the “BPLDA.” For G2 calculations, one obtains the “BGLDA” by the substituion $`gg(𝐫)`$ everywhere. Then one needs $`m_T(𝐫)`$ $`=`$ $`{\displaystyle \frac{d^2𝐩}{(2\pi )^3}u(𝐩,𝐫)v(𝐩,𝐫)[2n(𝐩,𝐫)+1]}`$ (39) $`=`$ $`g(𝐫)n_c(𝐫){\displaystyle \frac{d^3𝐩}{(2\pi )^3}\frac{1}{2ϵ}\left[2n(𝐩,𝐫)+1\right]}`$ (41) $`\mathrm{\Theta }\left((𝐩,𝐫)^2g^2n_c^2(𝐫)\right).`$ The integral is not formally convergent, however. Since the anomalous averages appear only in the context of the G1 and G2 approximations, where the formal ultraviolet divergence is eliminated, we may safely neglect the $`+1`$ term following the $`2n(𝐩,𝐫)`$. The semiclassical HFBP approximation exhibits a gapless excitation spectrum only if the condensate is also treated within the LDA, which implies the TF density: $`n_c(𝐫)`$ $`=`$ $`{\displaystyle \frac{\stackrel{~}{\mu }V_{\mathrm{ext}}(𝐫)2gn_T(𝐫)}{g}}`$ (43) $`\mathrm{\Theta }\left[\stackrel{~}{\mu }V_{\mathrm{ext}}(𝐫)2gn_T(𝐫)\right],`$ The TF approximation is valid in the limit of large $`N_c`$, where the energy contribution from the mean-field (Hartree) potential exceeds that of the kinetic energy. For this reason, Eq. (43) is not expected to be a good representation of the condensate density close to the transition temperature. In the regime of small condensate numbers, therefore, it becomes more important to solve the equations for the condensate and excitations exactly in order to obtain the low-lying discrete states, as described in the previous section. In this work, we use the exact GP and HFB equations, but the sum over discrete states is combined with an energy integral over high-lying states using LDA functions in the manner described by Reidl et al. : $$n_T(𝐫)=\underset{j}{}n_j(𝐫)\mathrm{\Theta }(ϵ_cϵ_j)+_{ϵ_c}^{\mathrm{}}𝑑ϵn_T(ϵ,𝐫),$$ (44) where $`n_j(𝐫)`$ is the $`j`$th term of Eq. (14), $`ϵ_c`$ is a low-energy cutoff, and, in the above notation, $`n_T(ϵ,𝐫)`$ has the form $`n_T(ϵ,𝐫)`$ $`=`$ $`{\displaystyle \frac{m^{3/2}}{2^{3/2}\pi }}\left[2n(𝐩,𝐫)+1{\displaystyle \frac{ϵ}{}}\right]`$ (45) $`\times `$ $`\left[V_{\mathrm{ext}}+\stackrel{~}{\mu }2gn\right]^{1/2}.`$ (46) A similar equation applies to the anomalous average $`m_T`$. This latter hybrid procedure is referred to below as the Discrete Quasiparticle Sum (DQS) approximation, an abbreviation for ‘discrete Hartree-Fock-Bogoliubov quasiparticle sum.’ Either a Popov or G2 approximation may be made within the DQS, and these are referred to below as DQSP and DQSG, respectively. A simpler LDA may be formulated by treating the local excitations within the Hartree-Fock approximation, which ignores the linear dispersion at low energies. The condensate density may again be obtained within the TF approximation using Eq. (43). The thermal density is given by $`n_T(𝐫)=\frac{d^3p}{(2\pi )^3}n(𝐩,𝐫)`$, where $`n(𝐩,𝐫)`$ is defined in Eq. (38) but with $`ϵ(𝐩,𝐫)=(𝐩,𝐫)`$. Integration over the momenta readily yields $$n_T(𝐫)=\frac{1}{\lambda _T^3}g_{3/2}\left(e^{\beta (V_{\text{ext}}(𝐫)+2gn(𝐫)\mu )}\right),$$ (47) where the thermal de Broglie wavelength is $`\lambda _T=(2\pi \mathrm{}^2/mkT)^{1/2}`$ and $`g_\alpha (z)=_{j=1}^{\mathrm{}}z^j/j^\alpha `$. As usual, the chemical potential $`\mu `$ is determined by the condition that the total atomic number, $`N=𝑑𝐫[n_c(𝐫)+n_T(𝐫)]`$. With the TF expression (43) for the condensate, the argument of the $`g_{3/2}`$ function in Eq. (47) is always less than unity. If an ‘exact’ solution for the condensate is used (i.e. obtained by solving the GP equation), the results are generally improved, but as noted below and in Ref. , there is then a range of temperatures $`TT_c`$ for which the $`g_{3/2}`$ function given in Eq. (47) diverges, since its argument can become greater than unity. An even simpler form of the LDA has been formulated in which the effect of interactions on the excited states is completely ignored. Assuming a TF form for the ground state, this LDA consists of the parametrically coupled equations (in view of the other approximations here, in these equations we ignore the distinction between $`\stackrel{~}{\mu }`$ and $`\mu `$): $`n_c(𝐫)={\displaystyle \frac{\stackrel{~}{\mu }V_{\mathrm{ext}}(𝐫)}{g}}\mathrm{\Theta }[\stackrel{~}{\mu }V_{\text{ext}}(𝐫)]`$ (48) $`n_T(𝐫)={\displaystyle \frac{1}{\lambda _T^3}}g_{3/2}\left(e^{\beta \left|V_{\text{ext}}(𝐫)\stackrel{~}{\mu }\right|}\right).`$ (49) In this approximation, the interaction enters only via the chemical potential in the TF equation, which is a function of $`a_{\mathrm{sc}}`$ and condensate number. For a spherical condensate, $`\stackrel{~}{\mu }_{\mathrm{TF}}=\frac{1}{/}2(15N_ca_{\mathrm{sc}}/a_0)^{2/5}\mathrm{}\omega _0`$, where $`a_0=\sqrt{\mathrm{}/M\omega _0}`$ is the bare oscillator length. It is shown in Ref. that a low order expansion of Eq. (49) yields the following expression for $`N_0/N`$: $$\frac{N_c}{N}=1\left(\frac{T}{T_c^0}\right)^3\eta \frac{\zeta (2)}{\zeta (3)}\left(\frac{T}{T_c^0}\right)^2\left(\frac{N_c}{N}\right)^{2/5},$$ (50) where $`\zeta (n)`$ is the Riemann zeta function, $`\eta =\stackrel{~}{\mu }_{\mathrm{TF}}/k_BT_c^0\frac{1}{/}2\zeta (3)^{1/3}(15N^{1/6}a_{\mathrm{sc}}/a_0)^{2/5}`$, and the critical temperature for $`N`$ ideal Bose atoms in a harmonic trap is given by $$k_BT_c^0/\mathrm{}\omega _0=0.9405N^{1/3}0.6842+0.50N^{1/3}$$ (51) Equation (50) is solved iteratively for $`N_c/N`$. ### E Ideal Bose gas Some of the plots given below contain results for ideal non-interacting Bose atoms ($`a_{\mathrm{sc}}=0`$) in a harmonic trap. The results given for $`N_c`$ were obtained from sums over the occupation numbers as given in Eq. (13), with $`d_j=2\mathrm{}_j+1`$, $`E_j=\mathrm{}\omega (\mathrm{}+2n_j+3/2)`$. The chemical potential $`\mu `$ was adjusted to satisfy the condition $`N=_{j=0}N_j`$. An alternative expression can be obtained from the density distribution given by Chou et al: $`N`$ $`=`$ $`{\displaystyle \frac{z_1}{1z_1}}`$ (52) $`+`$ $`{\displaystyle \underset{l=1}{\overset{\mathrm{}}{}}}z_1^l\{[(1e^{2l\beta })(\mathrm{tanh}(\beta l/2)]^{3/2}1\},`$ (53) where $`z_1=e^{\beta (\mu 3/2)}`$. This expression requires fewer terms than the aforementioned procedure, and gives identical results for temperatures up to about $`0.9T_c`$. ## III Computational techniques With a spherically symmetric trapping potential, all observables may be decomposed into functions of radius $`r`$ and spherical harmonics $`𝒴_l^m(\theta ,\varphi )`$. The GP and Bogoliubov equations then become one-dimensional in $`r`$; the ground state is assumed to have $`(\mathrm{},m)=(0,0)`$, while the excitations obtained using the Bogoliubov equations are $`2\mathrm{}+1`$ degenerate. Both equations are solved using the discrete variable representation (DVR), a computationally efficient approach for the trapped interacting Bose gases that has been recently described in detail . We have used two variants of the DVR approach: an equidistant mesh array derived from sine functions as discussed by Colbert and Miller , and a mesh based on Gaussian quadrature, using the zeros of associated Laguerre polynomials $`L_{N_L}^\alpha (r)`$, where $`N_L`$ is the order of the the quadrature and $`\alpha =2`$ for a spherical condensate . The latter DVR has the advantage of having a fine mesh for small $`r`$ where the condensate density is non-zero, and a more coarse mesh at larger distances where the thermal distribution varies smoothly. Though the condensate and excited orbitals are computed on the physical grid, the matrix elements of the operators are represented by Laguerre polynomials up to the order defining the Gauss quadrature $`N_L`$, which in the present calculations range from 1,000 to 2,800; matrix elements of the kinetic energy are computed from expressions given in Ref. . Increasing the value of $`N_L`$ increases the accuracy of high-lying states, allowing for a larger cutoff energy $`ϵ_c`$ at which the discrete sums are terminated, and a smaller number of atoms in the LDA integrals. Since high-order polynomials extend far beyond values of $`R_{\mathrm{max}}50a_0`$ relevant to trapped condensates, the number of spatial grid points required can be limited to just $`N_g200`$ for all values of $`N_L`$. Implementation of the above mean-field theory requires a stable and efficient iteration procedure to solve the GP and Bogoliubov equations for a given total number of atoms $`N`$ and temperature $`T`$. In our approach, the functions $`n_c(r)=\mathrm{\Phi }^2(r),n_T(r)`$, and $`m_T(r)`$ are calculated self-consistently using Eqs. (6) and (12)-(15), supplemented by the LDA expressions for states above the cutoff $`ϵ_c`$, for fixed $`N_c`$ and $`T`$; the chemical potential $`\mu `$ is determined by Eq. (27). Because this iteration procedure is especially delicate near $`T_c`$, yet is crucial for the results presented, we give a few more details. We emphasize that the convergence criterion must consider the spatial distribution functions $`n_c(𝐫)`$ and $`n_T(𝐫)`$ rather than simply the aggregate values, $`N_c`$ and $`N_T`$. The iterative procedure can be decomposed into three separate levels of self-consistency, subject to the minimization of the ‘Error’: $$\mathrm{Error}=𝑑𝐫\left[|n_c^{\mathrm{out}}(𝐫)n_c^{\mathrm{in}}(𝐫)|+|n_T^{\mathrm{out}}(𝐫)n_T^{\mathrm{in}}(𝐫)|\right].$$ (54) The ‘in’ and ‘out’ functions are the input and output of the combined GP and HFB equations plus the high energy LDA integral. Normally, the Error diminishes (though not necessarily monotonically) through level 1 iterations, in which the output functions are fed back into the GP, HFB and high energy LDA equations. In this level, the condensate number $`N_c`$ is held constant while the condensate density (normalized to unity) is allowed to vary. When the Error reaches some predetermined tolerance, level 2 iterations begin and $`N_c`$ is adjusted to approach the condition that $`N_c+N_T=N`$. The first level 2 adjustment from the converged level one iterations is based on a simple proportionality between $`N`$ and $`N_c`$. Subsequent level 2 adjustments are based on a linear relation between $`N_c`$ and $`N`$, where the parameters are obtained from the last two level 2 iterations. After $`N_c+N_T`$ has converged to $`N`$ to the desired tolerance, level 3 iterations proceed, in which iteration levels 1 and 2 are repeated with successively larger number of Laguerre functions $`N_L`$ and mesh points $`N_g`$. These three levels of iteration typically achieve accuracies for the condensate number $`N_c`$ of a few atoms. While this accuracy is beyond what is accessible to current experiments, it permits the comparison of different theoretical models. The iteration procedure is illustrated in Fig. 1, which tracks a calculation for $`210^5`$ atoms and scaled temperature $`t=k_BT/\mathrm{}\omega =53`$ (from Eq. (51), $`t_c^054.3`$), using the Laguerre DVR basis. After more than 50 iterations, $`N_c`$ converged from the initial estimate of 109 to the final value of 149 atoms (Fig. 1e). Each adjustment of $`N_c`$ (level 2) or $`N_L`$ (level 3) results in a jump in the error (Fig. 1f), which then converges again. In this calculation, $`N_L`$ increased from 1300 to 2100 (Fig. 1d), corresponding to an increase of mesh points (up to $`R_{\mathrm{max}}`$ = 42) from 149 to 190 (Fig. 1a), an increase in $`ϵ_c`$ from $`102\mathrm{}\omega _0`$ to $`144\mathrm{}\omega _0`$ (Fig. 1c), and a decrease in the fraction of total number of atoms in the LDA integral from 57% to 40% (Fig. 1b). The fraction of atoms in the LDA integral is negligible only for calculations at low temperatures with small $`N`$. Since $`T_c`$ rises as $`0.94N^{1/3}`$, the required number of thermal states rises with $`N`$ for calculations near $`T_c`$, and inevitably the LDA integration must include a larger fraction of atoms. For $`N=210^4`$, $`210^5`$m and $`10^6`$, at most 9%, 38%, and 74% of the atoms were in the integral at temperatures in the vicinity of $`T_c`$. Correspondingly, the mesh size $`N_g`$ required to ensure convergence increased from 140 to 210 for $`N`$ between $`10^3`$ and $`10^6`$. The reason $`N_g`$ does not increase more rapidly with $`N`$ is that the LDA approximation improves with the total number of atoms. It should be pointed out that for large values of $`N`$, the iteration procedure could exhibit instabilities when the temperature approached $`T_c`$. For $`N>10^5`$, we found that there often appeared to be (at least) two semi-stable regions when $`N_c5,000`$, between which the calculation tended to fluctuate. In order to ensure the solution remained in the more stable state, small temperature increments $`\mathrm{\Delta }t=0.2`$ were used. ## IV Thermal Averages ### A Condensate fraction In several of the plots to follow, results are presented for a series of values of $`a_{\mathrm{sc}}/a_0`$. For comparison with current experiments, we note that the scattering lengths $`a_{\mathrm{sc}}`$ for <sup>87</sup>Rb, <sup>23</sup>Na, and <sup>7</sup>Li are approximately given by $`110a_B`$, $`52a_B`$, and $`27.3a_B`$, respectively, where $`a_B5.29210^{11}`$ m is the Bohr radius. Thus, if one takes $`\omega =(\omega _x\omega _y\omega _z)^{1/3}`$, then for the recent MIT experiments on <sup>23</sup>Na, $`\nu =\omega /2\pi =96.4`$ Hz, the JILA experiments give $`\nu =182.5`$ Hz, and the Rice experiments give $`\nu =144.6`$ Hz, corresponding to $`a_{sc}/a_0=0.00129`$, $`0.00729`$, and $`0.00046`$, respectively. Figure 2 illustrates the consequences of setting the eigenvalue of the GP equation $`\stackrel{~}{\mu }`$ equal to the chemical potential $`\mu `$, as discussed in Section II A. With this assumption (here used in conjunction with the Popov approximation, $`m_T=0`$), $`N_c`$ goes to zero abruptly with $`T`$ when the population in excited states reaches the total number of atoms $`N=5,000`$. By contrast, results for $`a_{sc}=0`$, obtained as described in Sec. IIE, have a smooth tail at high temperature. Thus, in the limit $`a_{\mathrm{sc}}0`$, the results near $`T_c`$ exhibit a discontinuity with respect to the ideal gas results. Figures 3 and 4 show results obtained from calculations in which the chemical potential is as given in Eq. (27). The smooth variation of the chemical potential, Fig. 3, through $`T_c`$ is reflected in all relevant properties of the system, including the number of condensate atoms and excitation frequencies. When $`a_{\mathrm{sc}}>0`$, the chemical potential evolves continuously from positive to negative values, relative to the harmonic oscillator zero point energy $`\frac{3}{/}2\mathrm{}\omega `$, as the temperature increases. Since $`\mu `$ increases with the interaction strength, the value at which the chemical potential passes through zero increases with $`a_{\mathrm{sc}}`$ even though $`T_c`$ decreases. In addition, Fig. 3 shows that for $`a_{\mathrm{sc}}<0`$, $`\mu <3\mathrm{}\omega /2`$ everywhere, with maximum values at temperatures $`TT_c`$. Figure 4 shows the number of condensate atoms as a function of temperature for $`N=1,000`$ and $`20,000`$ for a range of interaction strengths $`a_{\mathrm{sc}}/a_0`$, calculated within the DQS formalism. The condensate population near $`T_c`$ is evidently a continuous function of both the scattering length and temperature. The plots shown in Fig. 4, especially for $`20,000`$ atoms, show that the G2 renormalization procedure results in a significantly higher value of $`N_c`$, relative to that obtained within the Popov approximation, for the larger values of $`a_{\mathrm{sc}}`$. Furthermore, the difference between the G2 and Popov results becomes more pronounced as $`a_{\mathrm{sc}}`$ increases. This behavior is consistent with expectation because G2 produces a weakening of the atom-atom interaction. The use of the occupation factors (25) rather than (13) also increases the value of $`N_c`$ by a few atoms at high temperatures, but the effect is much smaller than what results from the use of G2 theory. For $`a_{\mathrm{sc}}<0`$, the $`N_c`$ values reach a maximum when the calculation becomes numerically unstable , reflecting the physical instability of the cloud towards spatial collapse. The maximum $`N_c`$ values depend on $`a_{\mathrm{sc}}`$, as shown by the termination of the curves for these cases. For $`T=0`$, the maximum value is given by $`N_c^{\mathrm{max}}=0.573a_0/a_{sc}`$ . This critical number is known to decrease when $`T>0`$ due to the presence of thermal atoms . In these plots, the maximum $`N_c`$ is 80% to 57% of the value calculated for $`T=0`$, confirming that the thermal cloud significantly decreases the stability of the condensate for $`a_{\mathrm{sc}}<0`$. ### B Comparison with LDA and QMC It is interesting to explore how our finite temperature results compare with those obtained by other methods. Local density approximations are much simpler to implement numerically than the full self-consistent HFB equations and their variants. The opposite is true of Monte Carlo calculations, but these do not invoke the mean-field approximation so yield results for equilibrium configurations that are essentially exact. Fig. 5 for $`N=210^4`$ compares $`N_c`$ values from the Popov and G2 quasiparticle sums (DQSP and DQSG) with severak LDA methods. Our Hartree-Fock LDA (HFLDA) solves the GP equation for the condensate $`n_c(r)`$, iterated to self-consistency using Eq. (47) for the thermal distribution $`n_T(r)`$. We found it most efficient to start at low temperature, in order to obtain good initial estimates of $`n_T(r)`$ at successively higher values of $`T`$. No solution could be found for $`N_c/N<0.035`$ due to the failure of the HFLDA, as discussed above and in Ref. . The ‘semi-ideal’ LDA (SILDA) omits the $`n_T(r)`$ term in the TF expressions for the condensate (43) and for the total density $`n_T(r)`$ in Eq. (47). This results in the simple expressions (49) which are related solely through the chemical potential. Iterative solution of these equations yields results that are close to the other functions plotted in Fig. 5. The actual $`n_T(r)`$ distribution calculated with this approach exhibits a sharp peak at edge of the condensate due to the discontinuity at the Thomas-Fermi condensate radius. The inset of Fig. 5 shows that the Hartree-Fock Bogoliubov LDA methods, BPLDA and BGLDA, agree most closely with the hybrid method, DQSP and DQSG, respectively. The two BLDA methods employ a TF condensate, and thus the $`n_T(𝐫)`$ functions exhibit a small spike at the edge of the condensate, which has a cusp. As with the HFLDA, the calculations required iteration to self-consistency, which was facilitated when initial values were obtained by extrapolation from results from lower temperature values. It is remarkable that the values for $`N_c`$ from BLDA calculations agreed with the corresponding DQS results to better than 0.4% of $`N`$ in every case for which results were obtained. Even for HFLDA and SILDA, the differences with DQS results are less than the fractional error in current experiments. Thus these comparisons show that relatively simple LDA expressions are useful for obtaining the condensate fraction as a function of temperature. It is only in the region near $`T_c`$ and above, where the condensate number becomes small, that our LDA methods failed. The Quantum Monte Carlo (QMC) approach uses the exact Hamiltonian with a hard-sphere atom-atom interaction. Based on extensive numerical experience with <sup>4</sup>He , QMC should be most useful for the calculation of equilibrium quantities, such as the condensate fraction. Holzmann et al. have provided benchmark QMC calculations for the case of $`10^4`$ Bose atoms confined in a spherical trap, with $`a_{\mathrm{sc}}/a_0=0.0043`$. Table I shows comparisons between our results and those of QMC for the condensate number as a function of temperature. The DQSP, DQSG, and QMC values differ by up to 1.2% of the total atom number $`N`$. It is notable that at higher temperatures, $`N_c`$ falls off less quickly using HFBP and G2 than QMC. This may be due in part to the fact that the relationship between $`N_c`$ and $`\mu `$ in Eq. (27) is not entirely correct at higher temperatures, as discussed in Section II B, and may resemble ideal gas statistics too closely. Presumably the many-body effects that necessitate the renormalization of the atom-atom interaction are already included in the QMC procedure, in which case results with G2 should be closer than Popov to the QMC. Indeed, for $`t=k_BT/\mathrm{}\omega <17`$, the Popov results lie below QMC, while the G2 numbers are higher and closer to QMC. Above a scaled temperature $`t=18`$, however, the G2 results rise above QMC values. ### C Critical Temperature vs. $`a_{\mathrm{sc}}`$ Figures 4 and 5 show that large values of $`a_{\mathrm{sc}}`$ have the effect of flattening the curve of condensate number as a function of temperature, as is already apparent in the plots of Giorgini et al. . If these curves are fit to a function $`N_c/N=1(T/T_c)^\alpha `$, one obtains values for $`\alpha `$ as low as $`1.4`$, compared with the ideal gas value of $`3`$. Another parameter to characterize the effect of atom-atom interactions is the shift of the critical temperature from the ideal Bose gas case. For the homogeneous Bose gas, where it is uniquely defined as the point at which $`N_c`$ goes to zero, this shift has been the subject of intense discussion recently . For atoms in a harmonic potential, as is especially clear in Fig. 4, this point is not sharp (indeed, the number of condensate atoms is finite at all temperatures in a mesoscopic system). Definitions of $`T_c`$ that have been proposed include the point at which the density at the origin reaches the critical density for a homogeneous gas , the maximum of the specific heat, and the maximum of the temperature derivative of the specific heat . Since such energy-weighted properties pose additional problems for numerical calculations of thermal averages, $`T_c`$ is determined here as the maximum of the function $`d^2N_c/dT^2`$. The inflection point of the $`N_c`$ vs$`T`$ function, or zero of $`d^2N_c/dT^2`$ deviated from Eq. (51) by a significantly larger amount. Figure 6 shows $`T_c`$ values extracted from the data used in Fig. 4. For comparison, the ideal gas data are analyzed in a similar manner, yielding values of $`T_c`$ that are close to, but not identical with, those obtained using Eq. (51). Figures 6a and 6b correspond to $`1,000`$ and $`20,000`$ atoms, respectively. The inset in Fig. 6b shows how the transition temperature is determined from the data in a typical case, by making use of the three functions $`N_c(T)`$, $`dN_c(T)/dT`$, and $`d^2N_c/dT^2`$. Since both the condensate number and its temperature derivative are nearly straight lines, accurate calculation of the second derivative requires accurate numeric values of these functions. A semiclassical analysis by Giorgini et al. indicates that the transition temperature should decrease linearly from the ideal gas value with increasing particle interactions. The results of the DQS-Popov calculations confirm this general scaling; furthermore, as the number of atoms increases, the observed shift in the critical temperature $`\delta T_c`$ matches the semiclassical expression more closely at larger $`a_{\mathrm{sc}}/a_0`$. In contrast, with the DQS-G2 approach $`\delta T_c`$ shows significant deviations from linear scaling for small $`N`$, and these become more pronounced as the number of atoms increases. For $`N=210^4`$, the shifts are significantly less than the semiclassical values for the larger values of $`a_{\mathrm{sc}}/a_0`$ considered. ### D Renormalization of the atom-atom interaction As indicated in Fig. 4 above, the G2 renormalization yields values for $`N_c/N`$ that reflect the weakening of the atom-atom repulsion; at any given temperature, the number of atoms in the condensate increases relative to the value obtained using the Popov approximation. Perhaps more interesting is the spatial variation of the effective interaction in the harmonic trap. The renormalization is governed by the local value of $`m_T`$ relative to $`n_c`$. In general, $`|m_T|`$ increases with the number of noncondensed atoms $`n_T`$ since more terms enter the sum (15); however, $`m_T`$ vanishes when $`n_c=0`$, since the ‘quasihole’ amplitude $`v_i=0`$. In general, therefore, one might expect the local renormalized interaction to reach a minimum at some temperature. For a uniform Bose gas, this minimum occurs at exactly the transition temperature, and corresponds to a vanishing of the effective scattering length . In Fig. 7 we compare the condensate and thermal densities with the spatial variations of the anomalous average and the effective particle interactions for the case of $`N=20,000`$ and $`a_{\mathrm{sc}}/a_0=0.0072`$ for various temperatures. (As noted above, this would correspond to a relatively tight trap for <sup>23</sup>Na.) For these plots, $`\delta =0.01`$ in Eq. (34). There is a slight dependence of the results on $`\delta `$, since much smaller values $`\delta 10^4`$ lead to a small bump in the $`m_T(r)`$ function at the very edge of the condensate. The dependence on this arbitrary parameter indicates an ambiguity in the theory; however, the integrated numbers are not significantly altered by the choice of $`\delta `$, since the errors are incurred in regions of very small condensate density. The manner in which $`g(r)/g`$ attains a minimum in $`r`$ is shown in Fig. 8 for the particular case of $`N=210^5`$. The global minimum occurs at a temperature close to $`T_c`$, defined above. Following this procedure, we consider the $`g_{\mathrm{min}}(r)/g`$ functions for various values of $`N`$ for $`a_{\mathrm{sc}}/a_0=0.0072`$, which are displayed in Fig. 9. Though we have increased $`N`$ without changing the trap frequency, the approach to the thermodynamic limit is beginning to emerge. The minimum for each $`N`$ is found to always occur very close to the calculated transition temperature, and its value decreases approximately with $`\mathrm{log}(N)`$ over the range of $`N`$ considered. For $`N=10^6`$, we obtain $`g_{\mathrm{min}}(r)/g0.2`$. It should be noted that although the fraction of total atoms in the LDA integral increased to approximately $`\frac{3}{/}4`$ for $`N=10^6`$ near $`T_c`$, the high-energy LDA contribution to $`m_T`$ was in every case less than 2%, and typically an order of magnitude less than this value. It should be emphasized that the G2 renormalization employed in the present calculations is derived for a uniform Bose gas, and should best represent large condensate densities or low temperatures where the LDA is most applicable. While the LDA is bound to fail for $`TT_c`$, the regime where it loses validity will become smaller with increasing $`N`$, and should approach the critical region where perturbation theory itself breaks down. It would be preferable to define the renormalization of the particle interactions in terms of the full many-body T-matrix in a trap, and we hope to pursue this issue in future work. The G2 approach as formulated above, however, should properly describe the effects of two-body correlations for large trapped condensates at low to intermediate temperatures. Thus, the strong reduction in the effective interaction strength over much of the condensate, indicated by the G2 theory, could have significant experimental consequences. The predictions for the excitation frequencies are discussed further below. ### E Excitation frequencies The quasiparticle eigenvalues correspond to excitation frequencies, but it remains unclear what relationship exists between these values and experimentally observed resonances of the trapped gas at finite temperatures when the potential of a harmonic trap is perturbed periodically. In all mean-field calculations such as those presented here, the linear response equations assume that the thermal density is fixed, while in experiments it would also be perturbed. For this reason, the dipole excitation frequency obtained within mean-field theories will generally not satisfy the generalized Kohn theorem , which states that there is a mode in which the entire ensemble oscillates at the bare trap frequency. Calculations explicitly including the dynamics of both $`n_c`$ and $`n_T`$ are found to be consistent with the Kohn theorem. Figure 10 shows small but significant deviations in the Kohn mode from unity for $`N=210^4`$ and $`210^5`$, both within DQS-Popov and DQS-G2. That the G2 frequency should be lower than the Popov value cannot be simply understood in terms of an overall decrease in the interatomic repulsion, since this would predict a mode closer to unity. Rather, the spatial variation of the effective interaction leads to a flattening of the effective potential, comprised of the trap plus the Hartree potential; the looser effective confinement softens all the modes. We are not aware of other computational results in which the Popov value starts from below unity and rises above, before falling near $`T_c`$. This behavior may be a consequence of a more rigorous treatment of the chemical potential, Eq. (27). Alternatively, since the differences increase with $`N`$ (specifically, the non-condensate density), they may not have been observable with the smaller $`N`$ values studied previously. The temperature-dependence of the low-lying excitation frequencies obtained with the DQSP and DQSG approaches is shown in Fig. 11 for $`N=210^4`$ and $`210^5`$. The softening of all the excitation frequencies in the G2 approximation was found previously by the proponents of this theory (for a ‘pancake’ geometry) as well as by others using a similar perturbative approach to the interacting Bose gas . However, for a spherically symmetric trap, the results of Ref. for 2,000 Rb atoms showed only a negligible difference between Popov and G2 excitation frequencies. The present results show that for a spherically symmetric trap and larger atom numbers, there can be differences between the Popov and G2 values that would be experimentally detectable. These results also lead to the question whether for larger atom numbers, a renormalized atom-atom interaction would effect frequencies calculated by the methods of Refs. , which did not assume a static condensate. It should also be mentioned that experimentally observed excitation frequencies with larger numbers of sodium atoms in a ‘cigar’ geometry also exhibited a softening of both the quadrupole and dipole excitation frequencies as the temperature approaches $`T_c`$. ## V discussion and conclusions In this work, we have extended finite-temperature mean-field calculations for Bose-Einstein condensates confined in harmonic traps . A careful derivation of the mean-field equations provides improved definitions of the thermodynamic chemical potential and quasiparticle occupation factors, yielding observables that are continuous functions of the particle interactions. The numerical techniques employed in the calculations have allowed for the investigation of systems with the large numbers of atoms relevant to on-going experiments. In the process, we have been able to make several crucial comparisons between the results of evaluating discrete summations over quasiparticle states (which are numerically time-consuming) and various local density approximations. Furthermore, we have explored the implications of a recently proposed gapless theory which takes into account pairing correlations. The results presented above indicate a significant inadequacy of conventional static mean-field theory for computations of excitation frequencies of trapped Bose condensates at finite temperatures. For large number of atoms and interaction strength, we find appreciable deviations of the dipole frequency obtained with either the Popov or G2 approximations from expectations of the generalized Kohn’s theorem. In our computations, the condensate is static in the presence of thermal excitations. The excited dipole mode corresponds approximately to out-of-phase motion of the thermal cloud relative to the condensate, as observed experimentally when the dipole mode of the thermal cloud is excited separately. Detailed modelling of such excitation modes has been performed only by restrictive parametrization of the condensate and thermal cloud in the collisionless or hydrodynamic regimes. Both of these approaches address the two-fluid nature of these systems, and produce dipole modes that satisfy the Kohn theorem exactly. We will argue that equilibrium thermal excitations are computed accurately by the mean-field DQS methods presented here. However, any experimental probe of these excitations involves perturbative processes that require other theoretical methods. In principle, mean-field theories that include fluctuations in the population of excited states ought to be equivalent to the two-fluid dynamics in the collisionless regime. A full second-order perturbation theory of the interacting Bose gas should yield the coupled modes of the condensate and thermal clouds as well as damping rates. Indeed, employing the approximate many-body T-matrix in the calculations (the G2 approximation described above) yields excitations that have a temperature-dependence qualitatively similar to that of out-of-phase modes. We hope to explore these issues in future work. ###### Acknowledgements. This work was supported by the National Science Foundation (TB and BIS), the Office of Naval Research (TB and DLF), and by a grant from the NCF-Cray Foundation of the Netherlands (TB). The authors are grateful for valuable conversations with H. Beijerinck, K. Burnett, C. W. Clark, K. K. Das, M. Doery, M. Edwards, M. Gajda, A. Griffin, D. A. W. Hutchinson, H. Metcalf, H. Stoof, E. Vredenbregt, and E. Zaremba. The authors particularly appreciate S. A. Morgan for his valuable comments and for sending us a preliminary draft of his D. Phil. thesis.
warning/0001/math0001169.html
ar5iv
text
# Extremal metrics on graphs I ## Introduction There is a vast literature on the subject of extremal graph theory. There, the general approach is to consider a natural invariant (invariant with respect to isomorphism) of graphs, and to try to understand which graphs make the invariant as big as possible, subject to (presumably natural) constraints. Examples of such invariants are: Girth – the length of the shortest cycle; Diameter – the greatest distance between a pair of vertices; Tree number – the number of spanning trees, and some closely related spectral invariants: the “determinant of the Laplacian”, the smallest positive eigenvalue of the Laplacian, and so on. It is expected that graphs which are “good” with respect to any one of these invariant will be good with respect to the others (where by “good”, we mean that the graph is either extremal, or close to it), and will have other (a priori unsuspected) nice combinatorial properties. Extremal graph theory is a rather difficult subject, largely due to its intrinsically combinatorial nature (arguably it is this difficulty which attracts most of the practitioners). A seemingly not very closely related subject is that of differential geometry. One of its central areas is that of “uniformization”, or “optimal geometry”. There, we are often given a fixed topological space, and we try to find a metric on this space which maximizes some invariant. The actual invariants studied are very often similar to those mentioned above for graphs. The motivation, on the other hand, is sometimes the same as that of extremal graph theory, but sometimes there is an additional factor: it is hoped that the extremal metrics would give a canonical representation of the topological space, which renders its topological properties more transparent (for example, the study of the topology of the sphere would be much more difficult if we did not have its standard “round” representation at our disposal). Our motivation stems from both the areas sketched above: we would like to get canonical representations of graphs, but we have other concerns as well. First of all, the space of edge valuations of a given finite graph is a much simpler space than the space of metrics on a given topological space. Thus, we hope that the answers to our questions will be technically simpler than the corresponding differential-geometric results, but that the model is sufficiently rich to suggest what one might expect. By the same token, the space of edge valuations on a fixed graph is a much simpler space than the (discrete) space of graphs, though the latter is naturally embedded in the former. We thus hope to get insight into problems in extremal graph theory as well. ### 0.1. Outline of the paper We set up the basic deformation spaces and announce the main convexity results in Section 1. We set up the girth problem in Section 2, and characterize the extremal valuations in Section 4. We define the basic matrices and operators we are working with, and show the convexity of the bottom eigenvalue and the complexity in Section 3. We characterize valuations extremal for complexity (or “determinant of laplacian” in Section 5, and valuations extremal for the bottom eigenvalue in Section 6. Finally, in Section 7 we analyse completely those graphs which are extremal for $`\lambda _1`$, under the additional assumption (which turns out to be very strong), that $`\lambda _1`$ appears without multiplicity. ## 1. The foundations We will always consider a fixed finite simple graph $`G`$. We will consider the following deformation spaces of edge valuations on $`G`$: $$P(G)=\{f:E(G)^+|\underset{eE(G)}{}f(e)=|E(G)|\},$$ – the space of all edge valuations of $`G`$. $$T(G)=\{fP(G)|\underset{e_i\text{ incident to }v}{}=d_v,vV(G)\}$$ $$C(G)=\left\{fP(G)\right|g:V(G)^+,\text{such that}f(e_{vw})=g(v)+g(w)\}$$ The letters $`T`$ and $`C`$ are meant to suggest Teichmüller space and conformal deformation space respectively. All three spaces have a natural linear structure, which we will use without further comment. We will look at the variation of following invariants (defined below) over the above-described deformation spaces: girth $`g`$, bottom positive eigenvalue $`\lambda _1`$ of the Laplacian $`\mathrm{\Delta }`$ and $`\mathrm{log}det^{}\mathrm{\Delta }`$. The first striking observation about these invariants is the following: ###### Theorem 1. The quantities $`g`$, $`\lambda _1`$, $`\mathrm{log}det^{}\mathrm{\Delta }`$ are *convex* on $`P(G)`$ (and hence on its linear subspaces $`C(G)`$ and $`T(G)`$). (The proofs of these results are spread out through this paper: The convexity of girth is given in Section 2; the convexity of $`\mathrm{log}det^{}\mathrm{\Delta }`$ – by Theorem 4, and the convexity of $`\lambda _1`$ is outlined in Section 3.2.) ###### Remark 1. It can also be shown that the “topological entropy of the geodesic flow”, defined in terms of a different deformation of the adjacency matrix, is also convex. This is done in the article \[Riv99\] by the second author. The convexity has far-reaching consequences. To wit, for every invariant $`I\{g,\lambda _1,\mathrm{log}det^{}\mathrm{\Delta }\}`$, and for each deformation space $`D\{P,C,T\}`$ there is a unique canonical edge valuation $`G_D^I`$ maximizing the invariant. A natural question is one of the characterization of these critical valuations, and of understanding the relationship between the various $`G_D^I`$ for the different choices of $`I`$ and $`D`$. Some properties follow immediately from the convexity, in particular: ###### Observation 1. If $`G`$ possesses a group $`\mathrm{\Gamma }(G)`$ of automorphisms, then the weights of the critical points $`G_D^I`$ are invariant by these symmetries, thus, if the automorphism group of $`G`$ is edge transitive, then $`G_D^I`$ are all equal (independently of invariant and deformation space), and are given by the constant weighing on the edges. If the automorphism group of $`G`$ acts vertex-transitively, then $`G_C^I`$ is given by the constant weighing. ###### Remark 2. A large class of graphs the automorphism group of which is vertex- but not edge- transitive is given by the Cayley graphs of finite groups. For graphs not known a priori to be symmetric, the supposition that the unweighted graph $`G`$ is critical for one of the invariants, implies strong symmetry properties. For example, if $`G`$ is maximal for $`\mathrm{log}det^{}\mathrm{\Delta }`$, then there is the same number of spanning trees through every edge of $`G`$ ($`G`$ is equiarboreal in the terminology of Godsil). If $`G`$ is the maximum for $`\lambda _1`$ then, with rare exceptions, $`\lambda _1`$ occurs with multiplicity in the spectrum of $`\mathrm{\Delta }(G)`$. If $`G`$ is maximal for girth, then every edge of $`G`$ is contained in a shortest cycle (the precise somewhat stronger statement is the content of Theorem 7). ## 2. Girth The “direct” (girth) and “spectral” invariants are somewhat different conceptually. First we remind the reader that the length of a path in a weighted graph $`G`$ is the sum of the weights of the edges in the path. The girth $`\gamma (G)`$ is the length of the shortest cycle in $`G`$. The *distance* between two vertices of $`G`$ is the length of the shortest path connecting them; the diameter $`D(G)`$ is equal to the largest such distance. Thus, if $``$ is the set of all cycles of $`G`$, then the girth is given by: $$g=\underset{C}{\mathrm{min}}\underset{eC}{}f(e),$$ where $`f(e)`$ is the valuation of the edge $`e`$. Note that each of the terms $`_{eC}v(e)`$ is a linear function of the valuation $`f`$, and hence we have the immediate ###### Theorem 2. The girth $`g(f)`$ is a concave function on $`P(G)`$. ###### Proof. This follows from the observation that the minimum of a collection of concave (in particular linear) functions is concave. We leave the proof as an exercise to the interested reader. ∎ For two vertices $`u,v`$ of $`G`$ denote by $`\mathrm{\Lambda }(u,v)`$ the set of all paths $`\lambda `$ in $`G`$ connecting $`u`$ and $`v`$ (we can assume without loss of generality that the paths are not self-intersecting, to make sure that the number of paths considered is finite). Thus, $$d(u,v)=\underset{\lambda \mathrm{\Lambda }}{\mathrm{min}}\underset{e\lambda }{}f(e).$$ The diameter of $`G`$ is thus given by: $$D(G)=\underset{u,vV(G)}{\mathrm{max}}d(u,v).$$ Note that the diameter is *not* a priori convex, due to the additional maximum, though some of the methods we use for girth can be brought to bear on the diameter question as well. ## 3. Spectral invariants Let $`G`$ be a graph with $`n`$ vertices and $`m`$ edges (we denote the set of such graphs by $`𝒢_{m,n}`$), and $`f:E(G)`$ be a valuation on the edges of $`G`$ (in the sequel the valuations are always assumed positive, but this is not essential for the definitions below). The adjacency matrix $`A(G)`$ (or just $`A`$, when no ambiguity is possible) of a graph $`G𝒢_{m,n}`$ is a square matrix of size $`n`$ where $`A_{i,j}`$ is weight of the edge joining the vertices $`v_i`$ and $`v_j`$ if there such an edge, and $`0`$ otherwise. We always consider loopless graphs, so $`A_{ii}=0`$. The nearest neighbor Laplacian $`\mathrm{\Delta }`$ acts on functions on the set $`V(G)`$ of the vertices of $`G`$: given $`g:V(G)`$, $`\mathrm{\Delta }(g)(v)=_{wv}f(vw)(g(v)g(w))`$. (We exceptionally use $`vw`$ to denote the edge joining $`v`$ to $`w`$). Let $`\mathrm{\Delta }=\{L_{ij}\}`$ be the matrix of $`\mathrm{\Delta }(G)`$ of a (not necessarily simple) graph $`G`$; then $`L_{ii}`$ is the degree of the vertex $`v_i`$, and $`L_{ij}`$ is the number of the edges joining $`v_i`$ and $`v_jv_i`$ (equal to $`0`$ or $`1`$ for simple graphs). For $`k`$-regular graphs $`\mathrm{\Delta }=k\mathrm{Id}A`$. Let $`A`$ be the adjacency matrix of a weighted graph $`G`$ with $`n`$ vertices, let $`\delta `$ be the maximal degree of a vertex in $`G`$, and let its spectrum (in the decreasing order) be given by (1) $$\delta \mu _1>\mu _2\mu _3\mathrm{}\mu _n(\delta ).$$ The spectrum of $`\mathrm{\Delta }(G)`$ is $`0=\lambda _0<\lambda _1\mathrm{}\lambda _{n1}`$. For $`k`$-regular graphs, $`\delta =k=\mu _1`$ and $`\lambda _j=k\mu _{j+1},j=0,1,\mathrm{},n1`$. Note that the $`0`$ is always in the spectrum of $`\mathrm{\Delta }(G)`$ independently of the weighing on the edges, and, furthermore, as long as the weighing is strictly positive, and the graph $`G`$ is connected, the eigenspace of $`0`$ is spanned by the vector $`\mathrm{𝟏}=(1,\mathrm{},1).`$ ### 3.1. Complexity of a graph An important invariant of an *unweighed* graph $`G`$ is the number $`\tau (G)`$ of spanning trees of $`G`$; it is sometimes called the complexity of $`G`$. By Kirckhoff’s theorem (\[Kir\]), (2) $$n\tau (G)=\lambda _1\lambda _2\mathrm{}\lambda _{n1}$$ and $`\tau (G)`$ is equal to the determinant of any cofactor of the matrix $`\mathrm{\Delta }`$ and (it is some times called the determinant of Laplacian). The definitions for weighed graphs are the essentially the same, except that $$\tau (G)_f=\underset{T\text{spanning trees of }G}{}\underset{eE(T)}{}f(e).$$ This has a natural interpretation in the framework of electrical circuits, where $`f(e)`$ is thought of as the *conductance* of the edge $`e`$. See \[Bol98\]. ### 3.2. Variational problems The functions we consider are: The bottom nonzero eigenvalue $`\lambda _1`$ and $$\mathrm{log}det\mathrm{\Delta }^{}=\underset{i=1}{\overset{n1}{}}\mathrm{log}\lambda _i.$$ The bottom non-trivial eigenvalue $`\lambda _1`$ can be alternatively defined by the Rayleigh-Ritz quotient: (3) $$\lambda _1=\underset{_{i=1}^nx_i=0}{\mathrm{min}}\frac{x,\mathrm{\Delta }x}{x,x}$$ From this definition, the concavity of $`\lambda _1`$ over $`P(G)`$ is immediate. The concavity of $`\mathrm{log}det\mathrm{\Delta }^{}`$ is somewhat trickier. First we show: ###### Theorem 3. The logarithm of the determinant is a concave function on the set of positive definite symmetric matrices ###### Proof. Let $`Q`$ be such a matrix, and let $$Q(t)=Q+tB,t$$ be a line of symmetric matrices through $`Q`$. Then $$\frac{d\mathrm{log}det(Q(t))}{dt}=\mathrm{tr}(BQ^1),$$ and $$\frac{d^2\mathrm{log}det(Q(t))}{dt^2}=\mathrm{tr}(BQ^1BQ^1).$$ It suffices to show that the last trace is strictly positive. The matrix $`R=Q^1`$ is positive definite, so can be conjugated by an orthogonal matrix $`P`$ to a diagonal matrix $`D`$, where $`D_{ii}>0`$. So, we can rewrite $$\mathrm{tr}(BRBR)=\mathrm{tr}(BODO^tBODO^t)=\mathrm{tr}((O^tBO)D(O^tBO)D).$$ Let $`B^{}=O^tBO`$. $`B^{}`$ is still symmetric. We see that $`\mathrm{tr}(BRBR)=\mathrm{tr}(B^{}DB^{}D)`$. Now, let $`d`$ be the vector of the diagonal entries of $`D`$. It is not hard to check that $`\mathrm{tr}(B^{}DB^{}D)=d^td,`$ where $`_{ij}=b_{ij}^2.`$ Note, however, that by our assumptions, all the entries of $`d`$ are strictly positive, so $`d^td>0`$, and the result follows. ∎ Now we can prove ###### Theorem 4. The function $`\mathrm{log}det\mathrm{\Delta }^{}`$ is concave on $`P(G)`$. ###### Proof. The vector $`\mathrm{𝟏}=(1,\mathrm{},1)`$ is the zero eigenvector of $`\mathrm{\Delta }(G_f)`$ for any edge-valuation $`f`$ in $`P(G)`$. Thus, the restriction $`\mathrm{\Delta }^{}`$ of $`\mathrm{\Delta }`$ to the orthogonal complement of the subspace generated by $`\mathrm{𝟏}`$ is a symmetric positive-definite operator, whose entries as a matrix, furthermore, are obviously linear in those of $`\mathrm{\Delta }`$, no matter which basis of $`\mathrm{𝟏}^{}`$ we take. The result now follows immediately from Theorem 3 In the sequel, we characterize the extremal valuations for girth, $`\lambda _1,`$ and $`\mathrm{log}det\mathrm{\Delta }^{}`$ on our deformation spaces $`P,C,`$ and $`T`$. ## 4. Maximal girth valuations ### 4.1. Maximum in $`P(G)`$ Let $`G`$ be a fixed graph, and suppose that $`f_{\mathrm{max}}P(G)`$ is such that the $`g(f)`$ is maximal. There are two, somewhat different, cases to consider: the first is when $`f_{\mathrm{max}}`$ is an interior point of $`P(G)`$ (i.e. no $`f(e)`$ vanishes), the second is when $`f`$ is a boundary point (so that one for one or more edges $`e`$, $`f(e)=0`$). We will examine the interior point case first, since it contains the crucial ideas, and is slightly simpler. ### 4.2. Interior maximum The idea is that we use something like a piecewise-linear version of Lagrange multipliers. To wit, suppose that $`f_{\mathrm{max}}`$ is our maximal point. That means that there is a collection of cycles $`C_1,\mathrm{},C_n`$, such that $`\mathrm{}(C_i)=g_f,`$ while $`\mathrm{}(C)>g_f`$ for any other cycle $`C`$. Consider a small perturbation $`g`$ of the valuation $`f`$: $`f_1=f+th.`$ Since $`g`$ still has to lie in $`P(G)`$, we must have $`h,\mathrm{𝟏}=0.`$ The condition that $`f`$ is maximal is equivalent to saying that $`g_{f_1}g_f`$. However, for $`t`$ sufficiently small, a shortest cycle for the valuation $`f_1`$ has to be one of the cycles $`C_1,\mathrm{},C_n`$, thus the hypothesis that $`g_{f_1}g_f`$ means that $`\mathrm{min}\mathrm{}(C_i)`$ at the weighing $`f_1`$ has to be smaller than $`g_f`$. Consider the quantities $`H_i=_{eC_i}h(e)`$. We know that at least one of them has to be negative, but this (by multiplying by $`1`$ if necessary) is so if and only if $`i,j`$, such that $`sgnH_i=sgnH_j`$, or else all the $`H_i`$ vanish. The necessary and sufficient conditions follow from Farkas’ Lemma: ###### Theorem 5 (Farkas Lemma). Let $`v_1,\mathrm{},v_n,u^k`$. Then there exists a vector $`w^k`$, such that $`w,v_i0,1in`$ (at least one inner product being positive) and $`u,w=0`$ if and only if $`u`$ is *not* in the open convex cone generated by the $`v_i`$. Remark. $`u`$ is in the convex cone generated by the $`v_1,\mathrm{},v_n`$ if there exist $`\mu _1,\mathrm{},\mu _n`$ either all negative or all positive, such that $`u=_{i=1}^n\mu _iv_i`$. ###### Proof of Farkas Lemma. Suppose first that $$u=\underset{i=1}{\overset{n}{}}\mu _iv_i,\mu _i>0,1in.$$ Take any $`w`$ such that $`w,u=0`$. Then $$0=w,u=\underset{i=1}{\overset{n}{}}\mu _iw,v_i.$$ Since the $`\mu _i`$ are all positive, not all of the inner products $`w,v_i`$ can be positive, so $`w`$ does not satisfy the hypotheses of the theorem. Suppose now that $`u`$ is not in the open cone $`C`$ generated by $`v_1,\mathrm{},v_n`$. Consider the projection of $`C`$ onto the subspace $`u^{}`$ orthogonal to $`u`$. This is again an open convex cone $`C_u`$, which omits at least one point of $`u^{}`$ (the origin). Therefore it is a proper cone, and is thus contained in a half-space $`H^+`$, and thus the positive normal vector to $`H^+`$ has positive inner product with any vector in the projection of $`C`$, and hence with any vector in $`C`$ (since a vector in $`C`$ can be written as a sum of a vector in $`C_u`$ with a multiple of $`u`$). ∎ Theorem 5 can be generalized as follows: ###### Theorem 6. Let $`v_1,\mathrm{},v_n,u_1,\mathrm{},u_m^k`$. Then there exists a vector $`w^k`$, such that $`w,v_i0,1in`$ (with at one inner product positive) and $`u_j,w=0,1jm`$ if and only if *no* linear combination $`_{j=1}^ma_ju_j`$ is in the open convex cone generated by the $`v_i`$. ###### Proof. If some linear combination $`u=_{j=1}^mu_j`$ lies in the open cone $`C`$, then the same argument as in the beginning of the proof of Theorem 5 shows the non-existence of the requisite $`w`$. Otherwise, if $`u_1,\mathrm{},u_j`$ span $`^k`$, there is nothing left to prove. Assume then that they span a proper subspace $`U`$, and project $`C`$ onto the orthogonal complement, to get $`C_U`$. $`C_U`$ omits the origin by assumption, and the same argument as in the proof of Theorem 5 completes the proof. ∎ ###### Remark 3. The above theorems 5 and 6 do not address the question of when the there exists a nonzero vector such that the inner products with the $`u_j`$ *and* $`v_i`$ are all zero. This, however, is obviously true if and only if the span of all of the $`v_i`$ together with all of the $`u_j`$ is a proper subspace of $`^k`$. Theorems 5 and 6 and Remark 3 combine to give the following characterization of the extremal points of girth in $`P(G)`$, $`T(G)`$ and $`C(G)`$, which we state in the Theorem 7 below. First Notation. The systoles of $`G`$ corresponding to a weighing $`f`$ are cycles $`s_1,\mathrm{},s_k`$ whose length is equal to the girth of $`G`$ with the weighing $`f`$. We call *edge systoles* the vectors $`𝔰_1,\mathrm{},𝔰_k`$ in $`𝐑^{E(G)}`$ whose $`e`$-th coordinate is $`1`$ if $`e`$ is contained in the corresponding cycle $`s_j`$. We call the *vertex systole* corresponding to $`s_i`$, the vector $`\sigma _i`$ in $`^{V(G)}`$, whose $`v`$-th coordinate is $`1`$ if $`v`$ is incident to $`s_i`$, and $`0`$ otherwise. The *vertex vector* $`w_v`$ is the vector in $`^{E(G)}`$ whose $`e`$-th coordinate is $`0`$ unless $`e`$ is incident to the vertex $`v`$, in which case the coordinate is $`1`$. The *degree vector* $`d(G)`$ is the vector in $`^{V(G)}`$ whose $`v`$-th coordinate is the degree of the vertex $`v`$. ###### Theorem 7. A weighting $`fP(G)`$ is maximal for girth if and only if the constant vector $`\mathrm{𝟏}`$ lies in the open cone generated by the edge systoles of $`G`$ with the weighting $`f`$. The maximal weighing $`f`$ is unique if and only if the edge systoles of $`G`$ corresponding to the weighing $`f`$ together with the constant vector span the whole space $`^{E(G)}`$. A weighing $`f`$ in $`T(G)`$ is maximal for girth if and only if some linear combination of the vertex vectors $`w_1,\mathrm{},w_{V_G}`$ lies in the open cone generated by the edge systoles of $`G`$ with the weighing $`f`$. The maximal weighing $`f`$ is unique if and only if the edge systoles and the vertex vectors span $`^{E(G)}`$. A weighing $`f`$ in $`C(G)`$ is maximal for girth if and only if the degree vector $`d(G)`$ is contained in the open cone generated by the vertex systoles of $`G`$. The maximal weighing is unique if and only if the degree vector together with the vertex systoles span $`^{V(G)}.`$ ## 5. The tree number By the weighted version of Kirckhoff’s theorem (\[Bol98\]) (4) $$\tau (G)=\underset{T𝒯(G)}{}\underset{e_jT}{}x_j$$ where the sum is taken over the set $`𝒯(G)`$ of the spanning trees of $`G`$. We will find necessary and sufficient condition for a valuation $`f`$ to be the critical point for $`\tau (G)`$ (which is the same as being maximal by $`\mathrm{log}det\mathrm{\Delta }^{}`$, by the discussion in the Introduction) on $`P(G)`$, $`C(G)`$ and $`T(G)`$. It should be noted that such a critical point might not exist, and we might have to look for boundary maxima. Our methods can be easily adapted to deal with those cases as well, and since writing down the conditions is somewhat more cumbersome, we leave this to the reader. ### 5.1. Maximum in $`P(G)`$. We start with $`P(G)`$, since the result in that case is the simplest to state, and seems, at least at the moment to have the simplest combinatorial interpretation. Finding the maximum of $`\tau (G)`$ on $`P`$ is a Lagrange multiplier problem. The condition for $`xP(G)`$ to be a critical point for $`\tau (G)`$ is (5) $$\frac{\tau (G)}{x_1}=\frac{\tau (G)}{x_2}=\mathrm{}=\frac{\tau (G)}{x_m}$$ The partial derivatives above are given by $$\tau _j=\frac{\tau (G)}{x_j}=\underset{e_jT𝒯(G)}{}\underset{kj;e_kT}{}x_k.$$ The ratio $`\tau _j/\tau (G)`$ is called the effective resistance of $`e_j`$. We have thus proved: ###### Proposition 1. The graph valuation $`f`$ is maximal for $`\tau (G)`$ in $`P(G)`$ if and only if the effective resistances of all edges are the same. If an unweighted graph satisfies the assumptions of Proposition 1 then every edge of this graph is contained in the same number of spanning trees. Such graphs were studied by Godsil in \[God81\]; he calls these graphs equiarboreal. Obviously, all edge-transitive graphs (the automorphism group acts transitively on the edges) are equiarboreal.<sup>1</sup><sup>1</sup>1See \[Bou\] for examples of edge-transitive graphs which are not vertex-transitive. Godsil gives several more sufficient conditions for a graph to be equiarboreal; in particular, any distance-regular graph and any color class in an association scheme is equiarboreal (the least restrictive condition Godsil gives is for a graph to be 1-homogeneous). By an easy counting argument one can show that for an unweighted equiarboreal graph (6) $$T_1=T_2=\mathrm{}=\tau (G)(n1)/m,$$ where $`T_j`$ is the number of spanning trees containing $`e_j`$ (this is actually the result of Foster, cf. \[Fos\]) so the necessary condition for a graph to be equiarboreal is that $`m`$ divide $`(n1)\tau (G)`$. ###### Remark 4. Any tree is equiarboreal. We remark that the graphs which have the most spanning trees among the regular graphs with the same number of vertices are not necessarily equiarboreal, and vice versa. For example, the $`8`$-vertex Möbius wheel (cf. \[Big93\]) which has the most spanning trees among the $`8`$-vertex cubic graphs is not equiarboreal (cf. also \[Val\]), while the cube (which is certainly edge-transitive, hence equiarboreal) has the second biggest number of spanning trees among the $`8`$-vertex cubic graphs. ### 5.2. Maxima in $`T(G)`$ and $`C(G)`$ The Lagrange multiplier method of the previous section works just as well in $`T(G)`$ and $`C(G)`$. We leave the (easy) computation to the reader, and just summarize the results in ###### Theorem 8. A valuation $`f`$ is maximal in $`T(G)`$ if and only if there exists constants $`\lambda _1,\mathrm{},\lambda _{V(G)}`$, such that if the edge $`e`$ has endpoints $`v_i`$ and $`v_j`$, then $$\tau (e)=\lambda _i+\lambda _j.$$ A valuation $`f`$ is maximal in $`C(G)`$ if and only if for any two vertices $`v`$ and $`w`$ $$\mathrm{deg}w\underset{e\text{ incident to }v}{}\tau (e)=\mathrm{deg}v\underset{e\text{ incident to }w}{}\tau (e).$$ If we ask the same question as previously – when is the constant valuation maximal? – the condition for a maximum in $`T(G)`$ does not appear to have an obvious combinatorial interpretation. The condition for the maximum in $`C(G)`$ can be restated in the following way: ###### Corollary 1. Let $`d_T(v)=_{\text{spanning trees }T}\mathrm{deg}v\text{ in }T`$. Then, if the constant valuation is maximal for $`\tau (G)`$ on $`C(G)`$, then for any two vertices $`v`$ and $`w`$, $$\frac{d_T(v)}{\mathrm{deg}v}=\frac{d_T(w)}{\mathrm{deg}w}.$$ ## 6. Eigenvalues of the Laplacian To find the condition for maximality with respect to the bottom non-zero eigenvalue of the Laplacian, we will use the Rayleigh-Ritz characterization of of $`\lambda _1`$. This implies immediately that: ###### Theorem 9. Let $`f`$ be the weighing on $`G`$ (in our application, $`S`$ could be any one of $`P(G)`$, $`T(G)`$, $`C(G)`$, but it could be anything). Let $`E_{\lambda _1}`$ be the eigenspace corresponding to $`\lambda _1`$. Let $`Q`$ be any infinitesimal variation (that is, an element of the tangent space of $`S`$) of the valuation, and $`Q_\mathrm{\Delta }`$ the induced variation of the Laplacian matrix. Then the quadratic form given by $`Q_\mathrm{\Delta }`$ restricted to $`E_{\lambda _1}`$ is indefinite if and only if $`f`$ is maximal with respect to $`\lambda _1`$. ###### Proof. The argument is a version of that given in the beginning of section 4.2. We use the Rayleigh-Ritz quotient characterization (given in eq. 3). The space $`E_{\lambda _1}`$ is precisely the set of vectors where the minimum is attained, so at any unit vector $`xE_{\lambda _1}`$, $`x,(\mathrm{\Delta }+tQ_\mathrm{\Delta })x`$ is strictly greater than $`y,(\mathrm{\Delta }+tQ_\mathrm{\Delta })y`$ for $`y`$ a unit vector in $`E_{\lambda _1}`$, for $`t`$ sufficiently small. Thus, the first variation of $`\lambda _1`$ is given by the first variation of $`\lambda _1`$ restricted to $`E_{\lambda _1}`$, and that is given precisely by the restriction of the quadratic form given by $`Q_\mathrm{\Delta }`$. Now, if that were definite, we would be able to increase $`\lambda _1`$ by applying either the variation $`Q_\mathrm{\Delta }`$ or $`Q_\mathrm{\Delta }.`$ Note now that the space of all possible variations of the Laplacian induced by changes in the edge valuations has a natural linear structure (one can think of it as a subspace of the tangent space to symmetric matrices). Call that space $`V_{\text{var}}.`$ If $`MV_{\text{var}}`$, then $`x^tMx`$ can be thought of as a scalar product of $`M`$ with a vector $`P_x`$, whose $`ij`$-th coordinate is given by $`x_ix_j`$ (this is just the outer product of $`x`$ with itself, the letter $`P`$ is used to point out that when $`x`$ is a unit vector, $`P_x`$ is just the projection on the subspace generated by $`x`$). Let $`𝒫_{\lambda _1}=\{P_x|xE_{\lambda _1}\}`$. If $`S^{}`$ is the orthogonal complement to the tangent space of the deformation $`S`$, Theorem 6 (whose proof does not use the finiteness of the sets involved) gives us: ###### Theorem 10. A valuation $`f`$ is maximal in $`S`$ with respect to $`\lambda _1`$ if and only if the intersection of $`S^{}`$ with the open cone generated by $`𝒫_{\lambda _1}`$ is nonempty. What is the “open cone generated by $`𝒫_{\lambda _1}`$” ? It is an easy exercise to show that this is precisely the set of positive self-adjoint operators on $`E_{\lambda _1}`$ (that is, operators for which $`E_{\lambda _1}`$ is an invariant subspace ; which are positive on that subspace, and zero elsewhere). so Theorem 10 can be restated as: ###### Theorem 11. A valuation $`f`$ is maximal in $`S`$ with respect to $`\lambda _1`$ if and only if $`S^{}`$ contains a positive self-adjoint operator $`\mathrm{\Lambda }`$ on $`E_{\lambda _1}`$. ###### Corollary 2. If $`E_{\lambda _1}`$ is one-dimensional, then $`f`$ is maximal if and only if $`S^{}`$ is spanned by $`P_v`$, where $`v`$ is a unit eigenvector of $`\lambda _1`$. All the above might sound somewhat abstract, so let us now specialize to the the deformation spaces we have in mind. First, consider $`P(G)`$. In this case, it is easy to check that $$\frac{\mathrm{\Delta }}{f(e)}=Q_e,$$ where, if the endpoints of $`e`$ are $`v_i`$ and $`v_j`$, then $`Q_{ii}=Q_{jj}=1`$ ; $`Q_{ij}=Q_{ji}=1`$, and all of the other entries are $`0`$. The general variation of $`\mathrm{\Delta }`$ is given by $`Q_\alpha =_e\alpha _eQ_e,`$ and in order to stay in $`P(G)`$, we must have $`\alpha ,\mathrm{𝟏}=0.`$ It can be seen that the variation space $`S`$ of the Laplacians is spanned by the vectors $`Q_{e_0}Q_e`$, where $`e_0`$ is an arbitrary fixed edge. ### 6.1. The first eigenvalue $`\lambda _1`$ appears without multiplicity. If the eigenspace of $`\lambda _1`$ is one-dimensional, and the eigenvector is $`v`$, then by Corollary 2, $`v^t(Q_{e_0}Q_e)v=0`$, for all $`e`$. If $`e`$ is an edge with endpoints $`x`$ and $`y`$, then a calculation shows that $`v^tQ_ev=(v_xv_y)^2`$, and so for a maximal valuation, we must have (7) $$v_xv_y=\pm c$$ for any adjacent pair of vertices $`x,y`$. We study graphs which have an eigenvector satisfying the condition given by eq. (7) in section 7, but it is a priori clear that this condition is very rarely satisfied, and “usually” graphs maximal for $`\lambda _1`$ have a higher-dimensional first eigenspace. Curiously, the same holds for the (smaller) deformation spaces $`T(G)`$ and $`C(G)`$. Indeed, consider first $`T(G)`$. There, the deformation space of the Laplacians is spanned by matrices $`Q_{e_1}Q_{e_2}`$, where $`e_1`$ and $`e_2`$ have a vertex in common. Thus, the same computation as that leading to eq. (7) gives that the eigenvector of $`\lambda _1`$ for a critical graph must satisfy: (8) $$v_xv_y=\pm c_x,$$ for any adjacent pair of vertices $`x,y`$. A priori, this seems somewhat weaker than the condition (7) (since $`c_x`$ now depends on $`x`$), but in fact it is clear that for a connected graph $`G`$, it is equivalent ; the case of $`G`$ disconnected is different, but not particularly interesting. For $`C(G)`$, the deformation space of Laplacians is generated by the differences $`\mathrm{deg}wM_v\mathrm{deg}vM_w`$, where $`M_i`$ is the matrix whose $`ii`$-th entry is the degree of the $`i`$-th vertex; $`M_{ij}`$ is equal to $`1`$ if $`v_j`$ is incident to $`v_i`$, likewise $`M_{ji}`$, and all other $`M_{jk}`$ are equal to $`0`$. If $`v`$ is a vector, then $$v^tM_xv=\mathrm{deg}xv_x^22\underset{yx}{}v_xv_y=2v_x\underset{yx}{}(v_xv_y)\mathrm{deg}xv_x^2.$$ If $`v`$ is an eigenvector of $`G`$ with eigenvalue $`\lambda `$, then $`_{yx}(v_xv_y)=\lambda v_x,`$ and so $$v^tM_xv=(2\lambda \mathrm{deg}x)v_x^2.$$ From the equation $`v^t(\mathrm{deg}wM_v\mathrm{deg}vM_w)v=0`$, it follows that: (9) $$(\frac{2\lambda }{\mathrm{deg}x}1)v_x^2=(\frac{2\lambda }{\mathrm{deg}y}1)v_y^2.$$ In particular, note that when the graph $`G`$ is regular, it follows that (10) $$|v_x|=|v_y|,$$ for any two vertices $`x`$, $`y`$. Remark. It is not difficult to construct regular graphs which have an eigenvector satisfying eq. (10): any such graph is constructed by taking an $`l`$ regular bipartite graph, whose vertex set is the union of the sets $`R`$ of red vertices and $`B`$ of black vertices, and constructing $`k`$-regular graphs with vertex sets $`R`$ and $`B`$ respectively (then adjoining their edge sets to that of the original bipartite graph). Then the function which is $`1`$ on $`R`$ and $`1`$ on $`B`$ is an eigenvector with eigenvalue $`2l`$. It is much less clear that this can be done in such a way that $`2l`$ is the lowest eigenvalue. ### 6.2. The general case. When the eigenspace of $`\lambda _1`$ has dimension possibly greater than $`1`$, Theorem 11, compact with the finite-dimensional spectral theorem (that a positive self-adjoint operator can be diagonalized, with respect to an orthonormal basis, with positive weights) gives us the following extensions of the results of the previous subsection: ###### Theorem 12. In order for a valuation $`f`$ to be maximal for $`\lambda _1`$ with respect to $`P(G)`$, it is necessary and sufficient for there to be an orthogonal basis $`v_1,\mathrm{},v_d`$ of $`E_{\lambda _1}`$, and a collection of non-negative constants $`c_1,\mathrm{},c_d`$, not all zero, and a constant $`c>0`$ such that for any pair of adjacent vertices $`x,y`$ of $`G`$ (11) $$\underset{i=1}{\overset{d}{}}c_i(v_i(x)v_i(y))^2=c.$$ In order for $`f`$ to be maximal for $`\lambda _1`$ with respect to $`T(G)`$, the same condition (11) holds, assuming that $`G`$ is connected. In order for $`f`$ to be maximal for $`\lambda _1`$ with respect to $`C(G)`$, there must be constants as above, such that for any vertex $`x`$ of $`G`$, (12) $$(\frac{2\lambda }{\mathrm{deg}x}1)\underset{i=1}{\overset{d}{}}(c_iv_i^2(x))=c.$$ ###### Remark 5. The condition (11) gives an embedding of the edge set $`E(G)`$ into a $`(d1)`$-dimensional ellipsoid $$\underset{i=1}{\overset{d}{}}c_iz_i^2=c$$ by the “differentials” $`z_i=dv_i(e)=v_i(x)v_i(y)`$, where we have chosen an arbitrary orientation of the edge $`e=(xy)`$. The corresponding vertex condition gives, for a regular graph, a similar embedding of the vertex set $`V(G)`$ by the eigenvectors $`v_i`$. ## 7. Graphs with an eigenvector of constant gradient We now study connected graphs which admit an eigenvector $`f:V`$ satisfying (7) for some $`c0`$. If $`c=0`$ then $`f`$ is a multiple of a constant vector and so has eigenvalue zero which is a contradiction. If $`c0`$ then it is easy to see that the graph $`G`$ cannot have odd cycles and hence is bipartite. Namely, let $`u_1u_2\mathrm{}u_l`$ be a cycle. Then (putting $`u_l=u_0`$) $`_{i=1}^l(f(u_i)f(u_{i1}))=0`$. But each term in the sum is equal to $`\pm c`$, and since the number of terms in the sum is odd, they cannot add up to $`0.`$ We now want to study the unweighted $`k`$-regular graphs which have an eigenvector (corresponding to an eigenvalue $`\mu >0`$) satisfying (7) (without necessarily assuming that $`\mu `$ is simple). We shall rescale the eigenvector so that $`c=1`$ in (7). From (7) it follows that for each vertex $`u`$ the expression $`\mu f(u)`$ can only take one of the values $`k,k2,k4,\mathrm{},k+2,k`$. Consider first the vertex $`u_0`$ where $`f(u)`$ takes its maximal value $`a`$ (by changing the sign if necessary we can assume that $`a>0`$). It follows that $`f`$ takes value $`a1`$ on all the neighbors of $`u_0`$, hence $$a\mu =k$$ Next, consider any neighbor $`u_1`$ of $`u`$. The value of $`f`$ at any neighbor of $`u_1`$ can be either $`a`$ (let there be $`r_11`$ such neighbors; $`u_0`$ is one of them); or $`a2`$ (it follows that there are $`kr_1`$ such neighbors). From the definition of the Laplacian it follows that $$\mu (a1)=k2r_1$$ It follows from the last two formulas that (13) $$\mu =2r_1$$ where $`r_11`$ is a positive integer. If $`r_1=k`$, then $`\mu =2k`$ is the largest eigenvalue of $`\mathrm{\Delta }`$. We next define the level of a vertex $`u`$ to be equal to $`j`$ if $`f(u)=aj`$; we denote the set of all vertices of $`G`$ at level $`j`$ by $`G_j`$. It is easy to see that if $`uG_j`$ has $`r_j`$ neighbors where $`f`$ takes value $`aj+1`$ then $$\mu (aj)=k2r_j$$ It follows that $`r_j`$ is the same for all $`uG_j`$. Using (13) we see that $$r_1j=r_j$$ Consider now a “local minimum” $`uG_N`$. Then $`r_N=k`$, and we see that (14) $$r_1|k$$ Let $`n_j`$ denote the number of vertices in $`G_j`$. Counting the vertices connecting $`G_j`$and $`G_{j+1}`$ in two different ways, we see that for all $`0jN1`$, $$n_j(kr_j)=n_{j+1}r_{j+1}$$ Consider the case $`r_1=1,\mu =2`$. It follows from the previous calculations that $`r_j=j`$ and that $`N=k`$. Accordingly, $`n_j=n_0\left(\genfrac{}{}{0pt}{}{k}{j}\right)`$ and (15) $$|G|=2^kn_0$$ We next describe a class of graphs admitting an eigenvector of $`\mathrm{\Delta }`$ with $`\mu =2`$ satisfying (7). An obvious example of such a $`k`$-regular graph is the $`k`$-cube, and any such graph has the same number of vertices as a disjoint union of $`n_0`$ cubes by (15). Start now with such a union, choose the partition of the vertices of each cube into “levels” and take two edges $`u_1u_2`$ and $`u_3u_4`$ in two different cubes such that $`u_1,u_3`$ are both in level $`j`$ while $`u_2,u_4`$ are both in level $`j+1`$. If we perform an edge switch $$(u_1u_2),(u_3u_4)(u_1u_4),(u_3u_2)$$ then the number of the connected components of our graph will decrease while the eigenvector $`f`$ will remain an eigenvector with the same eigenvalue. Performing sequences of edge switches as described above, we obtain examples of connected graphs satisfying (7) and (15) for any $`n_0`$. Conversely, it is easy to show that starting from a graph satisfying (7) and (15) and having chosen a partition of its vertices into levels one can obtain $`n_0`$ disjoint $`k`$-cubes by performing a sequence of edge switches as above. We now want to consider the case when $`\mu =\mu _1`$ is the lowest eigenvalue of the Laplacian. The first remark is that then necessarily $`\mu k`$, and $`\mu =k`$ only if $`G=K_{k,k}`$. Next, we want to consider “small” $`k`$ for which $`k2\sqrt{k1}`$ (the “Ramanujan bound”) is less than $`2`$ (this happens for $`3k6`$). It then follows from the results of Alon (\[Nil\]) that the diameter of $`G`$ (and hence the number of vertices in $`G`$) is bounded above. ###### Proposition 2. For $`3k6`$ there are finitely many $`k`$-regular graphs for which the condition (7) is satisfied for an eigenvector of $`\mu _1`$. We next discuss graphs which have an eigenvector satisfying (7) with the eigenvalue $`\mu =2r_1>2`$. Recall that by (14) $`r_1|k`$. By counting the edges connecting the vertices in two consecutive levels one can show (as for $`\mu =2`$) that the number of vertices satisfies (16) $$|G|=2^{(k/r_1)}n_0$$ Also, since any vertex $`u_1G_1`$ has $`r_1`$ distinct neighbors in $`G_0`$, $$n_0r_1.$$ It is easy to construct examples of regular graphs which have eigenvectors with the eigenvalue $`\mu >2`$ satisfying (7); the construction is similar to that for $`\mu =2`$. We summarize the previous results: ###### Theorem 13. Let $`G`$ be a $`k`$-regular graph which has an eigenvector of $`\mathrm{\Delta }`$ with an eigenvalue $`\mu `$ satisfying (7). Then $`G`$ is bipartite, $`\mu =2l`$ is an even integer dividing $`2k`$, the number of vertices of $`G`$ is divisible by $`2^{(k/l)}`$, and for $`n_0l`$ there exist such graphs with $`n=2^{(k/l)}n_0`$ vertices.
warning/0001/hep-th0001159.html
ar5iv
text
# 1 The scalar fields of vector super-multiplets of D=5 theory parameterize a manifold that consists of different branches and due to the attractor equations point where the normal vector is parallel to a given constant vector 𝛼_𝐼 are “fixed-points” or extrema of the superpotential. The straight lines correspond to F=0 domain and shaded areas to 𝐹<0 domains. UPR-868-T CALT-68-2259 CITUSC/00-007 hep-th/0001159 Anti-deSitter Vacua of Gauged Supergravities with 8 Supercharges Klaus Behrndt<sup>a</sup><sup>1</sup><sup>1</sup>1e-mail: behrndt@theory.caltech.edu and Mirjam Cvetič<sup>b</sup><sup>2</sup><sup>2</sup>2e-mail: cvetic@cvetic.hep.upenn.edu <sup>a</sup> California Institute of Technology Pasadena, CA 91125 CIT-USC Center For Theoretical Physics University of Southern California Los Angeles, CA 90089-2536 <sup>b</sup> Department of Physics and Astronomy University of Pennsylvania, Philadelphia, PA 19104-6396 ## Abstract We investigate supersymmetric extrema of Abelian gauged supergravity theories with non-trivial vector multiplets and 8 supercharges in four and five dimensions. The scalar fields of these models parameterize a manifold consisting of disconnected branches and restricting to the case where this manifold has a non-singular metric we show that on every branch there can be at most one extremum, which is a local maximum (for $`W>0`$) or a minimum (for $`W<0`$) of the superpotential $`W`$. Therefore, these supergravity models do not allow for regular domain wall solutions interpolating between different extrema of the superpotential and the space-time transverse to the wall asymptotically always approaches the boundary of AdS (UV-fixed points in a dual field theory). There has been renewed interest in supergravity theories which allow for anti de Sitter (AdS) vacuum solutions. On one hand, the AdS/CFT correspondence implies that domain wall solutions encode the information on the renormalization group (RG) flow of (strongly coupled) super Yang Mills theories as discussed in . On the other hand, in the Randall-Sundrum scenario one considers a domain wall in a five-dimensional AdS space-time which allows for localization of gravity near the wall. In both cases, the gravitational effects in the domain wall backgrounds play an essential role. In the context of fundamental theory it is essential to consider supergravity theories with a non-trivial potential and demonstrate the existence of the domain wall solutions with the desired gravitational effects. The first examples of the superymmetric domain walls were found in N=1 D=4 supergravity theory (for a review see ). These solutions are static and interpolate between isolated supersymmetric extrema of the scalar potential. The gravitational properties of these domain walls crucially depend on the features of the superpotential and have been classified in . Unlike supergravity theories with four supercharges, e.g., N=1 D=4 supergravity, which have a rich structure of possible domain walls, the field theory embedding of possible domain wall solutions becomes highly constrained in supergravity theories with at least 8 supercharges, i.e. (N=1, D=5), (N=2, D=4) or (N=4, D=3) supergravity. In these cases the structure of the potential is related to gaugings of isometries of the scalar field manifold and thus it is much more restricted. There are the following different possibilities to gauge these supergravity models: (i) to gauge a subgroup of the $`SU(2)`$-R-symmetry, (ii) to gauge isometries of the vector moduli space or (iii) to gauge isometries of hyper-multiplet moduli space. In four dimensions the different cases have been reviewed in and the 5-d cases are discussed in . We will show that for the case (i), i.e. for Abelian gauged supergravity with eight supercharges, all extrema of the superpotential are disconnected as long as we restrict ourselves to scalar field manifolds with a non-singular metric. It is therefore impossible to construct regular domain wall solutions interpolating between different extrema. In addition, the space-time transverse to the wall always approaches the boundary of AdS asymptotically and thus these solutions disallow for the localization of gravity near the interior of the wall, i.e. in D=5 it is impossible to embed the Randall-Sundrum scenario in this framework. <sup>3</sup><sup>3</sup>3The same conclusion seems to hold also for the case (ii). On the other hand, we restrict ourselves to studying vector multiplets only and will not consider case (iii). 1) $`D=5`$ Case For this case equivalent conclusions have been derived in . In D=5 supergravity with 8 supercharges the (real) scalars in the vector multiplets $`\varphi ^A`$ parameterize a hypersurface $``$ defined by a cubic equation : $$F(X)\frac{1}{6}C_{IJK}X^IX^JX^K=1$$ (1) where $`I=0,1,2,\mathrm{}n`$ and $`n`$ is the number of vector multiplets and $`C_{IJK}`$ are the coefficient defining the cubic Chern-Simons term in the supergravity Lagrangian. (In Calabi-Yau compactifications these are the topological intersection numbers.) In general $``$ is not connected and consists of different branches, separated by regions where $`F(X)<0`$; see figure 1. Gauging a $`U(1)`$ subgroup of the $`SU(2)`$ R-symmetry , the scalars remain uncharged, however they obtain a potential given by $$V=6\left(\frac{3}{4}g^{AB}_AW_BWW^2\right),$$ (2) with the superpotential and the metric: $$W=\alpha _IX^I,g_{AB}=\frac{1}{2}\left(_AX^I_BX^J_I_JF(X)\right)|_{F=1},$$ (3) From the M/string-theory perspective this superpotential appears due to calibrated sub-manifolds of the internal space, where the vector $`\alpha _I`$ corresponds to non-trivial fluxes (see ). On the other hand, from D=5 perspective this superpotential arises solely due to constraints of supersymmetry . Supersymmetric extrema of $`V`$ are given by extrema of $`W`$ and because $`_AX^I`$ defines tangent vectors on $``$, supersymmetric extrema are points on $``$ where the constant vector $`\alpha _I`$ is normal to $``$, i.e. where $$\alpha _IX_I,$$ (4) with $`X_I=\frac{1}{3}_IF(X)|_{F=1}`$ (see also figure 1). This is the essence of the attractor equation as derived in , which has been discussed in the domain wall context in . In addition, the second derivative of $`W`$ satisfies the following constraint : $$_A_BW=\frac{2}{3}g_{AB}W+T_{ABC}g^{CE}_EW,$$ (5) where $`T_{ABC}_AX^I_BX^J_CX^KC_{IJK}`$. At the extrema of the superpotential (fixed-points) $`_EW=0`$ and thus eq. (5) implies that for a positive definite scalar metric $`g_{AB}`$ these extrema can only be minima (for $`W>0`$) or maxima (for $`W<0`$), but not saddle points. Moreover this equation implies that the supersymmetric extrema always correspond to the maxima of the potential, i.e., $`_A_BV=4g_{AB}W^2`$, see also . Such two extrema, (see figure 2) could be connected if one allows for a saddle point in-between, however since we restricted ourselves to the physical domain of the scalar metric, i.e. we assume that $`g_{AB}`$ is positive definite, such saddle points are excluded and all the extrema of $`W`$ lie on disconnected branches of $``$<sup>4</sup><sup>4</sup>4 One arrives at the same conclusion if one assumes that $``$ is convex, which implies that there can only be one point on any given branch where (4) holds . Let us also point out that the space-time transverse to the wall necessarily asymptotes to the boundary of AdS and not the Cauchy AdS horizon, thus exceeding the one which is necessary for implementation of the Randall-Sundrum set-up. (Equivalent arguments are given in .) Namely, for the static domain wall Ansatz: $$ds^2=𝒜(z)(dt^2+dx_i^2)+dz^2$$ (6) the Killing spinor equations which fixes the scalars $`\varphi ^A(z)`$ take the form : $$_z\varphi ^A=\pm 3g^{AB}_BW,_z\mathrm{log}𝒜=2W,$$ (7) with the spinor constraint $`\mathrm{\Gamma }_zϵ=\pm ϵ`$. The expansion of the kink solution $`\varphi ^A=\varphi _{|\pm }^A+\delta \varphi ^A`$ around the supersymmetric extremum ($`_BW_{|\pm }=0`$) renders (7) in the following asymptotic form: $$_z(\mathrm{log}\delta \varphi ^A)=\pm 2W_{|\pm },_z(\mathrm{log}𝒜)=2W_{|\pm }.$$ (8) (In the derivation of the first eq. in (8) we employed (5), evaluated at $`_BW=0`$.) The first eq. in (8) implies that for a kink solution to approach (exponentially fast) the asymptotic values $`\varphi _\pm ^A`$ (as $`z\pm \mathrm{}`$) the superpotential $`W`$ has to satisfy: $`\mathrm{signW}_+=\mathrm{signW}_{}`$. As a consequence, the second eq. in (8) implies that in this case the metric coefficient $`𝒜`$ necessarily grows exponentially fast on either side of the wall. Thus, these walls, in addition to being singular, necessarily approach the boundary of the AdS space-time as $`z\pm \mathrm{}`$; they have a repulsive gravity on either side of the wall and thus cannot localize gravity. In a dual field theory these supersymmetric extrema always correspond to ultra-violet (UV) fixed-points . The one-scalar example of such walls were given in (see also for an early work on supergravity kinks); in the interior these walls have a power-law curvature singularity. Another comment is in order. If one does not insist on the positive definite scalar metric $`g_{AB}`$, some supersymmetric extrema can become saddle points<sup>5</sup><sup>5</sup>5For special parameter choices the extrema of $`W`$ may be at the boundary of $``$ and may not correspond to AdS vacua. We thank S. Gubser for communications on this point.. In this case it is possible to connect, e.g., a supersymmetric maximum with a supersymmetric saddle point in a continuous manner. However, again due to (8), the space-time on either side of such non-singular walls asymptotes to the boundaries of AdS, which are UV fixed points of the dual field theories. 2) $`D=4`$ Case As the second example we consider D=4 , N=2 gauged supergravity (for a review see ). In contrast to the D=5 case before, the scalars of vector supermultiplets are now complex and the potential is given by $$V=e^K\left(g^{A\overline{B}}D_AWD_{\overline{B}}\overline{W}3|W|^2\right),$$ (9) where $`W`$ is the superpotential, $`K`$ is the Kähler potential, $`D_AW\left(_A+(_AK)\right)W`$, and $`g_{A\overline{B}}=_A_{\overline{B}}K`$ is the Kähler metric. In comparison with 5-d supergravity we have to replace the scalars $`X^I`$ by the symplectic section $`(X^I,F_I)`$ where $`F_I=_IF(X)`$ denotes is derivative of the prepotential $`F(X)`$. The superpotential is again a linear function, but now in the symplectic section : $$W=\alpha _IX^I\beta ^IF_I.$$ (10) Supersymmetric extrema of $`V`$ are given by extrema of $`W`$ with respect to the covarient derivatives, i.e. $`D_AW=0`$. In order to facilitate the investigation of supersymmetric extrema we write the potential in terms of a real function $`\widehat{W}`$: $$\widehat{W}\xi |W|e^{K/2}=\xi |\alpha _IL^I\beta ^IM_I|,$$ (11) which is invariant under Kähler transformations and the analogous constraint to (1) becomes $`i(\overline{L}^IM_IL^I\overline{M}_I)=1`$. Here $`\xi =\pm 1`$ and can only change sign iff $`W`$ passes through zero. In terms of $`\widehat{W}`$, the potential (9) takes the form: $$V=3\left(\frac{4}{3}g^{A\overline{B}}_A\widehat{W}_{\overline{B}}\widehat{W}\widehat{W}^2\right).$$ (12) Since $`\widehat{W}`$ satisfies the relation: $`(_A\widehat{W})\widehat{W}^1=(D_AW)(2W)^1`$, extrema of $`\widehat{W}`$ correspond to the supersymmetric extrema of the potential. Note, that employing the real function $`\widehat{W}`$ the potential (12) has been cast in a form completely parallel to that of D=5 potential. In order to obtain the second derivatives at extrema, we can employ basic formulae from special geometry. Namely, the symplectic section $`𝒱=(L^I,M_I)=e^{K/2}(X^I,F_I)`$ satisfies $$D_AD_B𝒱=iC_{ABC}g^{C\overline{E}}D_{\overline{E}}\overline{𝒱},D_AD_{\overline{B}}𝒱=g_{A\overline{B}}𝒱,$$ (13) where $`C_{ABC}`$ is the covariantly holomorphic section. Moreover, using the definition of $`\widehat{W}`$ one finds for the second derivatives: $`(_A_B\widehat{W})\widehat{W}^1=(D_AD_BW)(2W)^1+𝒪(D_AW)`$ and because $`D_A𝒱(_A+\frac{1}{2}(_AK))𝒱=e^{K/2}D_AW`$ we obtain $`(_A_B\widehat{W})\widehat{W}^1=(D_AD_B𝒱)(2𝒱)^1+𝒪(D_AW)`$. As a consequence of (13) we find that at supersymmetric extrema ($`_A\widehat{W}=0`$) the second derivatives of $`\widehat{W}`$ satisfy: $$_A_{\overline{B}}\widehat{W}=\frac{1}{2}g_{A\overline{B}}\widehat{W},_A_B\widehat{W}=0.$$ (14) The relationships (14) imply the same conclusions as in D=5 case: for the domain with a positive definite Kähler metric $`g_{A\overline{B}}`$, i.e. restricting to the physical region of the metric, all the extrema of $`\widehat{W}`$ are either minima (for $`\widehat{W}>0`$) or maxima (for $`\widehat{W}<0`$), but never saddle points. This result again implies that the supersymmetric extrema are disconnected and that the potential always has maxima there, i.e. $`_A_{\overline{B}}V=g_{A\overline{B}}\widehat{W}^2`$ and $`_A_BV=0`$. For the purpose of addressing the space-time properties of supersymmetric (static) domain wall backgrounds one arrives at the following Killing spinor equations which we cast in an explicitly Kähler invariant form: $$_z\varphi ^A=2g^{A\overline{B}}_{\overline{B}}\widehat{W},_z\mathrm{log}𝒜=\widehat{W}.$$ (15) In addition, the complex scalar fields have to statisfy the “geodesic equation”: $$\mathrm{Im}\left[(_z\varphi ^A)_A(\mathrm{log}\widehat{W})\right]=0,$$ (16) while the Killing spinors satisfy: $`ϵ_\alpha =\xi i\gamma ^ze^{\theta _W}\epsilon _{\alpha \beta }ϵ^\beta `$, where $`\theta _W`$ is the phase of the holomorphic superpotential $`W`$. The geodesic eq. (16) is a supergravity generalization of the equation in global supersymmetric theory where a kink solution corresponds to a straight line in the W-plane ($`_z\theta _W=0`$) (see, e.g., ). The above Killing spinor equations were first derived for domain walls in D=4 N=1 supergravity , where the superpotential $`W`$ and Kähler potential $`K`$ are not subject to constraints of N=2 special geometry; of course for the N=1 case the equations remain the same, but with constrained $`K`$ and $`W`$. The expansion of the kink solution $`\varphi ^A=\varphi _{|\pm }^A+\delta \varphi ^A`$ around the supersymmetric extremum ($`_B\widehat{W}_{|\pm }=0`$) renders (15) in the following asymptotic form: $$_z(\mathrm{log}\delta \varphi ^A)=2\widehat{W}_\pm ,_z(\mathrm{log}𝒜)=2\widehat{W}_\pm .$$ (17) In the derivation of the first eq. in (17) we used the relationships (14). The first eq. in (17) implies that that for the existence of a kink solution ($`\varphi ^A\varphi _{|\pm }^A`$ as $`z\pm \mathrm{}`$) the superpotential $`W`$ necessarily crosses zero and thus $`\xi _{|+}=\xi _|`$. The second eq. in (17) in turn implies that in this case the metric coefficient $`𝒜`$ necessarily grows exponentially fast on either side of the wall. Thus, just as in the D=5 case, these walls are necessarily singular (because extrema are on different branches) and the space-time asymptotically approaching the boundary of the AdS on either side of the wall (UV fixed points). On the other hand, just as in the D=5 case, if one relaxes the constraint of positive definite Kähler metric, such domains could connect across a smooth region, but the asymptotic space-time remains to approach the boundary of AdS asymptotically. We have not considered the D=3 case with 8 supercharges, where the corresponding scalars parameterize a quarternionic manifold; we expect, however, the same conclusions. On the other hand, just as for the D=4 cases, breaking further supersymmetry (to four or only two supercharges) one expects a much richer structure (see ). Let us end with some general remarks. In our arguments it was important to assume that the Kähler metric is everywhere positive definite, which excluded saddle points and disconnected all supersymmetric extrema. This is a very strong restriction, which may not be the case for physically interesting applications. E.g., the manifold $``$ can have boundaries where eigenvalues of the Kähler metric vanish and additional massless modes are expected. In addition, we restricted ourselves to supersymmtric cases only, but it may be that the supersymmetric vacua are connected by non-BPS sphaleron configurations as recently discussed in . These are very interesting aspects, which certainly deserve further investigations. Acknowledgments The work is supported by a DFG Heisenberg grant (K.B.), in part by the Department of Energy under grant number DE-FG03-92-ER 40701 (K.B.), DOE-FG02-95ER40893 (M.C.) and the University of Pennsylvania Research Foundation (M.C). M.C. would like to thank Caltech High Energy Theory Group for hospitality during the completion of the work.
warning/0001/cond-mat0001075.html
ar5iv
text
# 1 Introduction ## 1 Introduction Measurement of the electromagnetic response of materials at microwave frequencies is important for both fundamental and practical reasons. The complex conductivity of metals and superconductors gives insights into the physics of the quasiparticle excitations and collective charge properties of the material. Interesting collective behavior which can be investigated through the complex conductivity includes superconductivity, spin and charge density waves, Josephson plasmons, etc. The dielectric properties of materials give insights into the polarization dynamics of insulators and ferroelectrics. A class of interesting insulating materials, the parent compounds of cuprate superconductors, combine both dielectric and antiferromagnetic behavior because of strong correlations between the charge carriers. Ferromagnetic materials interact with microwaves in several unusual ways, including ferromagnetic resonance and anti-resonance, and spin-wave resonance. In addition, a series of discoveries over the past several decades have found many materials which display coexistence of superconductivity with either antiferromagnetism, or modified forms of ferromagnetism. This wide range of interesting electromagnetic behavior of contemporary materials requires that experimentalists working in this field master many diverse measurement techniques and have a broad understanding of condensed matter physics. On the practical side, electromagnetic measurements are essential for creating new device technologies and optimizing existing devices and processes. For example, nonlinearities in superconducting materials limit their applications at microwave frequencies to relatively low power uses. Although intrinsic nonlinearities of superconductors will eventually limit their utility, most practical nonlinearities are caused by extrinsic defects in the material or device. Careful characterization of these materials on the length scale of the extrinsic inhomogeneities is required to tackle this problem. Another practical issue is the optimization of materials for frequency-agile applications. Here the issue is measurement of dielectric or magnetic properties which can be tuned with an electric or magnetic field, often in thin-film materials. Yet another important application of electromagnetic measurements is in diagnostic measurements. For example, in a semiconductor integrated circuit process, one needs to measure and control quantities such as sheet resistance, doping profiles, and electromigration in wires. As the speed of microprocessors continues to climb, the microwave properties of materials become increasingly important. Finally, if high-speed integrated circuits fail, it is important to locate the fault (often an open or short) as accurately as possible, and to correct the problem quickly to maintain a high yield. Electromagnetic measurements are thus an important foundation for many emerging technologies. In this paper we present a new paradigm for electromagnetic measurements of materials. We first briefly review the traditional methods of microwave measurements and point out some of their important limitations. We then present an alternative approach to these measurements through quantitative near-field microwave microscopy. The remainder of the paper is devoted to reviewing the progress we have made in this exciting new field of research. ## 2 Traditional Microwave Measurements of Electromagnetic Properties The fundamental electrodynamic quantities of greatest interest are the surface impedance Z<sub>s</sub>, the conductivity $`\sigma `$, the dielectric permittivity $`\epsilon `$, and the magnetic permeability $`\mu `$. A review of the definitions of conductivity and surface impedance of normal metals and superconductors is given elsewhere in this book. All of these quantities are complex and in general are a function of many variables, including frequency, temperature, magnetic field, electric field, etc. Traditional microwave measurements of these quantities are typically done on length scales of the free-space wavelength of the signal, which is about 3 cm at 10 GHz. The earliest microwave measurements on superconductors by Pippard were done with the sample acting as a quarter-wavelength resonator . In this case the electromagnetic properties measured are actually an average of the properties along the sample, weighted by the form of the standing-wave resonance pattern. In the case of elemental Pb or Sn this is not much of an issue, but in the case of complicated multi-element oxides it can give rise to misleading results. A refinement of this technique came through the use of cavity perturbation methods. In this case the sample is immersed in a larger electromagnetic cavity in a region of uniform electric or magnetic field . The properties of the sample can be deduced by comparing the resonant frequency shift and quality factor change from a well-characterized initial (unperturbed) state of the cavity. In this case one measures an average of the electromagnetic properties of the sample, again weighted by the distribution of fields and currents created in the sample . These field and current distributions can be quite complicated for most practical samples , becoming simple only for carefully shaped ellipsoids of revolution. Other resonant techniques have been developed in recent years, particularly for examination of thin-film superconducting materials. These include the parallel-plate resonator technique, in which two congruent films form a transmission line resonator with a length scale on the order of the microwave wavelength . Dielectric resonators are also sensitive to thin-film surface impedance but are averaged over an area on the order of the wavelength . Far-field surface-impedance microscopy techniques use a diffraction-limited beam in a confocal geometry to locally measure the surface impedance at millimeter-wave frequencies, but have a spatial resolution limited to a few millimeters . Non-resonant methods can also be employed to measure electrodynamic properties. For instance, waveguide transmission through thin films has been successful at determining the absolute value of the penetration depth, but provides an average of the properties over the film, again on the length scale of the microwave wavelength. ### 2.1 Limitations of Traditional Techniques Although traditional electrodynamics measurement techniques have been very successful, they do suffer from some fundamental limitations. First, they tend to measure a weighted average over large areas of the sample, on the scale of millimeters to centimeters. In the case of oxide thin films, it is often difficult to prepare a material which is homogeneous on such long length scales. For oxide single crystals, doping inhomogeneities have been shown to exist on many length scales from the nanometer to the millimeter range . In addition, surface morphology and second phases (such as flux particles) complicate the response of the crystal to electromagnetic radiation. Hence, traditional electrodynamics measurements yield only an average or overall picture of the sample properties. This means that the intrinsic behavior of the material can be masked by the response of a relatively small fraction of extrinsic or second-phase material. When new materials are measured, there is often a lingering doubt whether the response is due to new intrinsic physics or some unforeseen extrinsic effect. A second limitation of traditional techniques is the generation of large screening currents, especially near edges, and in the case of superconductors, the subsequent admission of rf magnetic flux into the sample. The rf flux will significantly increase the surface impedance and can easily dominate the response of the material. Such issues have arisen many times in the exploration of electromagnetic properties of new superconducting materials. For instance, early measurements of the penetration depth in thin films of YBCO were corrupted by vortex entry into the films ; harmonic generation in superconducting crystals can be dominated by edge effects ; and the nonlinear Meissner effect in superconductors can be swamped easily by vortex entry problems . ### 2.2 A new paradigm for electrodynamics measurements To overcome the limitations of traditional electrodynamics measurements, we have developed a new family of near-field microwave microscopes which permit local quantitative measurements of surface impedance and conductivity, as well as dielectric and magnetic properties. Near-field techniques also relieve us from the “light-cone constraint”, in which the length scales that we can probe with electromagnetic radiation are dictated by the frequency. For instance, this permits broadband electrodynamics measurements to be performed on very fine length scales. Because the spatial resolution is determined by geometrical features which we can control, a class of altogether new electromagnetic experiments can be performed. For example, we now can measure the local tunability of high-frequency electrical properties and make new connections between microstructure and its associated physical properties. ## 3 An Overview of Microwave Microscopy Techniques Near-field microwave microscopy is an art which has formed gradually over the years from many different sources. In principle the intellectual founder of the near-field microscopy effort was Synge in his prescient paper of 1928 . However, most practitioners of the art are not familiar with this seminal work despite its importance. Many near-field techniques were developed empirically, some, in fact, developed to solve practical problems in manufacturing. The earliest effort to perform high resolution quantitative microwave measurements seems to be from the ferromagnetic resonance community, led by the work of Frait , and later Soohoo . However, not all subsequent developments can be traced back to these roots. To codify the great variety of work on near-field microwave microscopy of materials, we present a simple picture of the five basic types of microwave microscopy in Fig. 1. Although these classifications are somewhat gross, they allow us to cleanly distinguish the main efforts in the field. Fig. 1(a) illustrates a traditional microwave cavity resonator with a small hole in one wall. The sample is placed in close proximity to the wall; a small region of the sample, defined by the hole diameter, perturbs the resonant frequency (causing a frequency shift $`\mathrm{\Delta }`$f) and quality factor (Q) of the resonator. Because the hole is so small (0.5 mm diameter for Frait ) and the sample makes such a small perturbation to the cavity, one must examine a highly lossy property of the sample, such as ferromagnetic resonance. Ferromagnetic resonance (FMR) gives information about the local internal fields and magnetization of the sample. Frait , Soohoo and Bhagat have successfully used this microscope for FMR imaging, while Ikeya has used it for electron spin resonance imaging. An important variation on this theme was developed by Ash and Nichols . They used an open hemispherical resonator with a flat-plate reflector. The plate had a small hole in it and the sample was positioned close to this hole, but outside the resonator. They did not study FMR, so in order to recover a perturbation signal from the resonator, the sample separation from the hole was modulated at a fixed frequency. They were able to phase-sensitively recover the reflected signal from the resonator to construct a qualitative image of the sample as it was scanned under the hole. These cavity methods all make use of an evanescent mode to couple the resonant cavity to a local part of the sample. In that sense these methods resemble the tapered optical fiber with aperture method of near-field scanning optical microscopy (NSOM) . Fig. 1(b) illustrates a class of non-resonant microscopes in which the sample is placed on or near the end of a microwave transmission line. The complex reflectivity R or transmission coefficient T is measured, and properties of the sample are deduced. The most common technique is to measure reflectivity from a coaxial transmission line terminated (possibly with an air gap) by the sample . Variations include the use of waveguides , waveguides covered by a resonant slit and microstrip lines in reflection. Transmission measurements also have been done in the coaxial and waveguide geometries. These techniques have mainly been used to map metallic conductivity or sheet resistance, and dielectric constant. In some cases the measurements are done quantitatively. Fig. 1(c) shows schematically one of the most sensitive forms of near-field microwave microscopy. The idea is that the sample is put near the open end of a transmission-line resonator, and changes in the resonant frequency and quality factor are monitored as the sample is scanned. This class of techniques differs from that shown in Fig. 1(a) through the use of a “field enhancing” feature at the end of the transmission line, rather than an evanescent aperture in the resonator. The field-enhancing feature sets the scale for the spatial resolution of this class of microscopes. The first embodiment of this concept was conceived for measuring the moisture content of paper . Other embodiments use coaxial transmission lines with the sample in contact with the open end , or with an air gap between the probe and the sample . A related far-field technique has been used to image surface resistance and nonlinearity with a scanned dielectric resonator in contact with the sample . The resonant transmission-line technique has proven to be the most quantitative of all the near-field microwave microscopy methods. Quantitative imaging of topography, sheet resistance, dielectric constant and loss, and other quantities has been achieved. In addition, the use of field enhancing features has pushed the spatial resolution of resonant and non-resonant microscopes to the sub-$`\mu `$m domain while maintaining the quantitative nature of the measurements. The last two classes of near-field microscopy techniques are less developed, but show great promise for the future. Fig. 1(d) shows a cantilever with a sharp tip positioned over the sample. This geometry can be used to perform atomic force microscopy to determine the topography of the sample, in addition to microwave microscopy. There are three sub-classes of scanning probe measurements of materials properties at microwave frequencies. The first is to perform localized electron-spin resonance measurements . The tip is used for STM, and a magnetic field is applied to the sample. It is found that a local signal at the Larmor precession frequency can be extracted from the tip and converted into an image of ESR response of the sample. A second embodiment is to create a magnetic field gradient on the sample (e.g., with a small magnetic particle on the tip) while immersing it in an rf magnetic field. The sample will locally satisfy the magnetic resonance condition and exert a force on the cantilever . A third scanning probe method is to simply use a sharp metallized tip to perform “apertureless” near-field microscopy. The sharp tip in close proximity to a metallic sample will locally enhance radiation introduced by a focused far-field beam. If an additional phenomenon can take place in this localized region due to the enhanced field strength, and it can be detected, one has a local microscopic probe of the physics associated with that phenomenon. One important example of this kind of microscopy is apertureless infrared microscopy, where the reflected signal from the region of the probe tip is measured while the tip is periodically dithered up and down . Fig. 1(e) shows the scanning superconducting quantum interference device (SQUID) method of microwave microscopy. The SQUID will generate circulating rf currents when a dc bias is placed across the loop. The frequency of these currents is directly proportional to the applied bias voltage. The currents generate rf magnetic fields which then impinge on the sample. The sample will generate its own response currents which in turn modify the inductance of the SQUID loop. By monitoring the magnetic-field feedback signal required to keep the SQUID in a constant flux state, one can map the electromagnetic response of the sample . This method has the advantage of being very broadband (in principle from rf up to the gap frequency of the superconductor used to make the SQUID, $``$100 GHz or more) and quantitative. ### 3.1 Detailed Description of the Maryland Microscope Our group has developed a novel form of near-field scanning microwave microscopy based on the scanned resonator technique (Fig. 1(c)). As shown in Fig. 2, our microscope consists of a resonant coaxial cable which is weakly coupled to a microwave generator on one end through a decoupling capacitor C<sub>d</sub>, and coupled to a sample through an open-ended coaxial probe on the other end. In the absence of a sample, the microscope is a half-wave resonator (Fig. 4). As a metallic sample approaches the open end of the coaxial probe, the boundary condition there goes progressively from open circuit to short circuit (e.g., for the blunt probe shown in Fig. 3(a)). In the limit of the sample touching the coaxial probe, the microscope becomes a quarter-wave resonator (Fig. 4). In the process, the resonant frequency shifts by one half the spacing between neighboring modes, which is on the order of 100 MHz. Hence the microscope frequency shift is very sensitive to the probe/sample separation, as well as the electrical properties of the sample (i.e., the conductivity, dielectric constant and magnetic permeability). Consider again the limit of the sample very far from the probe. In this case the quality factor of the resonator, Q, is high since only internal dissipation processes (and radiation losses) are active. As the sample approaches the open end of the coaxial probe it must be considered part of the resonant circuit. As such, it will add losses to the system and in general reduce the Q of the microscope. The Q thus gives insight into the additional loss mechanisms introduced by the sample. As the sample is scanned beneath the probe, the probe-sample separation will vary (depending upon the topography of the sample), causing the capacitive coupling to the sample, C<sub>x</sub>, to vary. This has the result of changing the resonant frequency of the coaxial-cable resonator. Also, as the local electrical properties of the sample vary, so will the resonant frequency and quality factor, Q, of the resonant cable. The goal then is to turn the measured quantities $`\mathrm{\Delta }`$f and Q versus position over the sample into quantitative images of sample electrical properties. A circuit is used to force the microwave generator to follow a single resonant mode of the cable, and a second circuit is used to measure the Q of the circuit, both in real time (Fig. 2) . The microwave source is frequency modulated by an external oscillator at a rate $`f_{FM}`$ 3 kHz. The electric field at the probe tip is perturbed by the region of the sample beneath the probe’s center conductor. We monitor these perturbations using a diode detector which produces a voltage signal proportional to the power reflected from the resonator. A feedback circuit (Fig. 2) keeps the microwave source locked to a resonant frequency of the transmission line, and has a voltage output which is proportional to shifts in the resonant frequency due to the sample. Hence, as the sample is scanned below the open-ended coaxial probe, the frequency shift and Q signals are collected by a computer. The circuit runs quickly enough to accurately record at scan speeds of up to 25 mm/sec. To determine the quality factor Q of the resonant circuit, a lock-in amplifier, referenced at $`2f_{FM}`$, gives an output voltage $`V_{2f_{FM}}`$ which is related to the curvature of the reflected power vs. frequency curve on resonance, and hence to Q. To relate $`V_{2f_{FM}}`$ and Q, we perform a separate experiment, in which we vary Q using a microwave absorber at various heights below the probe tip, and measure the absolute reflection coefficient $`\left|\rho \right|^2`$ of the resonator. If $`\left|\rho _0\right|^2`$ is the reflection coefficient at a resonant frequency $`f_0`$, then the coupling coefficient between the source and the resonator is $`\beta =\left(1\left|\rho _0\right|\right)/\left(1+\left|\rho _0\right|\right).`$ The loaded quality factor of the resonator is Q<sub>L</sub> $`=f_0/\mathrm{\Delta }f`$, where $`\mathrm{\Delta }f`$ is the difference in frequency between the two points where $`\left|\rho \right|^2=`$ $`\left(1+\beta ^2\right)/\left(1+\beta \right)^2`$. The unloaded quality factor, which is the Q of the resonator without coupling to the microwave source and detector, is then Q$`{}_{0}{}^{}=`$ Q$`{}_{L}{}^{}(1+\beta )`$. We also measure $`V_{2f_{FM}}`$, and find that there is a unique functional relationship between Q<sub>0</sub> and $`V_{2f_{FM}}`$; thus, we need to calibrate this relationship only once for a given microscope resonance. In a typical scan, we record $`V_{2f_{FM}}`$, and afterward convert $`V_{2f_{FM}}`$ to Q<sub>0</sub>. ## 4 Properties of the Near-Field Microwave Microscope ### 4.1 Spatial Resolution The spatial resolution of the microscope has been demonstrated to be the larger of the probe-sample separation and the diameter of the inner conductor wire in the open-ended coaxial cable . As with other forms of near-field microscopy, the probe is placed well within one wavelength of the sample under study. This is particularly easy to accomplish at rf, microwave and millimeter-wave frequencies because the wavelength ranges from meters to millimeters. However, the high spatial resolution comes from a “field-enhancement” feature of the probe. It has been found that sharp tips produce higher spatial resolution than blunt probes. Figure 3 contrasts these two types of probes in a transmission-line microscope, like those shown in Fig. 1(b) and (c). Figure 3(a) shows a blunt electric field probe, while Fig. 3(b) shows an STM-tipped probe. The blunt probe has a spatial resolution given by the area of the center conductor when operated in non-contact mode at a height smaller than the inner conductor diameter. The STM tip has a spatial resolution in contact mode on the order of 1 $`\mu `$m, depending somewhat on the bluntness of the tip. The “lightning-rod effect” is responsible for the increased spatial resolution with a sharp-tip probe. This same principle is used in apertureless near-field scanning optical microscopy. The basic idea is that the electromagnetic fields in the near-field environment can be treated in the static approximation. In electrostatic equilibrium there will be a large electric field at sharp corners, with the local field being inversely proportional to the radius of curvature. This enhancement means that the response of the sample immediately below the sharp tip will dominate the signal. Quantitative calculations of dielectric response from an STM-tipped probe bear out this qualitative picture . In addition to achieving high spatial resolution it also is important to vary the spatial resolution while maintaining quantitative measurement capability. There are times when a gross characterization or measurement of a sample property is required, for instance on the scale of a wafer or thin-film. There are physical property features on many length scales in materials. Hence, a microscope must adapt itself to imaging all of these scales. Our microscope is well adapted for this purpose and has been used with spatial resolutions of 500 $`\mu `$m, 200 $`\mu `$m, 100 $`\mu `$m, 12 $`\mu `$m and 1 $`\mu `$m. #### Evanescence and Microwave Microscopy An evanescent wave is one which does not propagate, but is exponentially attenuated with distance. Such behavior can be created with microwaves in a variety of situations. The simplest is to use a single-conductor cylindrical wave guide and to reduce its lateral dimensions such that the cutoff frequency is made greater than the frequency of the propagating mode. The wave is then attenuated on the length scale of the diameter of the waveguide . This is the method employed by the microscopes in Fig. 1(a). The attractive feature of evanescence is the fact that the wave equation for the electromagnetic fields reduces to Laplace’s equation in regions small compared to the wavelength of the radiation. This allows one to use low-frequency analysis to understand the distribution of the fields in the evanescent near-field region. Evanescence through an aperture brings with it several distinct disadvantages. First, evanescent waveguides are characterized by a purely reactive impedance , hence they reflect signals back to the source very efficiently. This leads to the problem of a very poor signal-to-noise ratio, as is encountered in traditional NSOM through tapered optical fibers. The near-field optical community is now moving towards the “apertureless NSOM” method in which no evanescent waves are used. Instead they rely on a field-concentrating feature to enhance the signal from a very localized part of the sample. This is similar to the methods which we employ in our microscopes. Another disadvantage of evanescence through an apeture at microwave frequencies is the long length scale of the evanescent decay. Evanescence is of great utility in scanning tunneling microscopy because the decay length is so short that it permits atomic resolution imaging. However, for radiation coming out of a hole beyond cutoff, the decay length scale is on the order of the hole diameter. This can be made no smaller than a few hundred microns in practical microwave microscopes without reducing the signal-to-noise ratio to unity. However, techniques of field enhancement can improve the spatial resolution to much smaller values without compromising the signal-to-noise. Here it may be that short-length-scale evanescent waves associated with the field enhancement feature may dramatically improve the spatial resolution . These evanescent waves are commonly encountered at the boundary between two different waveguides . The details of how these evanescent waves contribute to the imaging of electromagnetic properties on short length scales remain to be determined. ### 4.2 Contrast mechanisms The microwave microscope generates contrast for many different electromagnetic properties of materials. These include topography, conductivity, dielectric constant and magnetic permeability. In addition, other properties such as ferromagnetic and antiferromagnetic resonance produce contrast for the microscope. In order to quantitatively extract this information from the microscope images, it is necessary to model the microscope and its interaction with the sample. Here we begin with a model of our microscope, and defer the discussion of sample-specific modeling to the next section. The heart of the microscope is the microwave resonator (Fig. 5). It is coupled to the generator by the capacitor C<sub>d</sub> and to the sample through the capacitor C<sub>x</sub> (in the case of an electric-field probe). The sample can be modeled as an impedance Z<sub>x</sub> which includes resistive, capacitive and inductive loads which it presents to the microscope. Also included in the model are the directional coupler and detector which are used to measure the reflected signal from the resonator. One can calculate the signal measured in the detector V<sub>drc</sub> in terms of the source driving voltage, V<sub>source</sub>. In the simplest case, one finds the following expression for the reflectivity of the resonator: $$V_{drc}=\frac{Z_3Z_0}{Z_3Z_0}P(1P)V_{source},$$ (1) where Z$`{}_{3}{}^{}=Z_d+Z_0\frac{(Z_x+Z_0)e^{\gamma L}+(Z_xZ_0)e^{\gamma L}}{(Z_x+Z_0)e^{\gamma L}(Z_xZ_0)e^{\gamma L}},`$ Z<sub>0</sub> is the characteristic impedance of the transmission line, P is the directional coupler voltage fraction, Z<sub>d</sub> = 1/$`i\omega C_d`$, $`\gamma `$ is the complex propagation constant of the transmission line of length L, and Z<sub>x</sub> is the impedance of the capacitor C<sub>x</sub> in series with the sample. The presence of the sample modifies the resonance of the microscope in several ways. First, as the probe-sample distance varies, so does the capacitance C<sub>x</sub>. Consider a metallic sample below a blunt probe (e.g., Fig. 3(a)). If the sample is much closer than the probe diameter, there is a parallel-plate capacitance between the probe and the sample. As the sample approaches the probe, this capacitance will increase, and the resonant frequency will decrease. This is because the increased capacitance at the end of the transmission line will effectively lengthen the resonator. Ultimately as the sample comes into contact with the probe, a short-circuit condition is achieved and the microscope becomes a quarter-wave resonator. Hence, the greatest frequency shift expected is that from converting a half-wave to a quarter-wave resonator, which is half the distance between neighboring modes in either the open circuit or short-circuit microscope. ### 4.3 Frequency Coverage Because the microscope is a transmission-line resonator with many closely spaced modes, and because each mode can be used for imaging, the microscope is a tremendously broadband instrument. In a typical design, the transmission-line resonator has a length of approximately 1 m, giving a fundamental mode at about 125 MHz, and harmonic modes at integer multiples of this frequency. Hence, imaging can be done up to frequencies where higher-order modes begin to propagate in the coaxial cable, which can be over 100 GHz for small diameter cable. The only other limitation is the microwave electronics required to operate the microscope, such as the source, directional coupler, detector and connectors. However, using modern sources, electronics and the 1-mm standard connector, a 100+ GHz imaging bandwidth can be achieved. ### 4.4 Temperature range Our microscopes work well at room temperature. However, we also have constructed a cryogenic version of the microwave microscope. In this case only the probe and part of the resonator is actually held at cryogenic temperatures, along with the sample. Many images have been acquired at liquid-nitrogen temperature (77 K) in both materials and device-imaging modes . We see no reason why this cannot be extended to pumped helium temperatures (1.2 K). At the other extreme, high-temperature coaxial cables exist which can be used for imaging up to 1000C . ### 4.5 Other Independent Parameters One can imagine introducing other quantities into the sample during microwave microscopy. For instance, light illumination is an obvious degree of freedom. We have found that measurements of tunability are very interesting with the scanning near-field microwave microscope. In this case the use of external electric and magnetic fields is most important. #### Electric field We have developed a technique to introduce a local and tunable electric field in our microwave microscope when imaging in contact mode (Fig. 2). A dc electric bias voltage can be applied to the center conductor of the microwave probe through a bias tee. This will create an electric field between the probe tip and the surrounding grounded surfaces. By introducing a suitable ground plane, controlled electric fields can be applied to the sample being imaged at microwave frequencies . #### Magnetic field We also can introduce a static magnetic field on samples being imaged by the microwave microscope using an electromagnet. The field can be modulated for imaging magnetic resonant phenomena in the sample. ## 5 Quantitative Imaging with the Near-Field Microwave Microscope Here we present some of the results on quantitative imaging with our microscope. Very simply, one expects lossy properties of the sample shown in the model of Fig. 5 to affect the Q of the microscope, while reactive properties and topography will primarily affect the resonant frequency. ### 5.1 Topography Images obtained from resonant near-field scanning microwave microscopes are the result of two distinct contrast mechanisms: shifts in the resonant frequency due to electrical coupling between the probe and the sample, and changes in the quality factor Q of the resonator due to losses in the sample . Such images will inevitably contain intrinsic information about a sample (such as dielectric constant or surface resistance) as well as extrinsic information (such as surface topography). To facilitate quantitative imaging of intrinsic material properties, it is essential to be able to accurately account for the effects of finite probe-sample separation and topography. The interaction between the probe and a metallic sample can be represented by a capacitance C<sub>x</sub> between the sample and the inner conductor of the probe, analogous to the mechanism in scanning capacitance microscopy . For a highly conducting sample and a small probe-sample capacitance C<sub>x</sub>, i.e., C<sub>x</sub> $``$ L/cZ<sub>0</sub>, the reflected voltage V<sub>drc</sub> at the output of the directional coupler can be written as above in Eq. (1) with $`Z_3=Z_d+\frac{Z_0\left[\mathrm{cosh}\left(\gamma L\right)+i\omega C_xZ_0\mathrm{sinh}\left(\gamma L\right)\right]}{\mathrm{sinh}\left(\gamma L\right)+i\omega C_xZ_0\mathrm{cosh}\left(\gamma L\right)},`$ where c is the speed of light, and $`\omega `$ is the angular frequency of the source. A plot of Eq. (1) versus frequency would show a series of dips corresponding to enhanced absorption at the resonant frequencies . For C<sub>x</sub>=0 and C<sub>d</sub>=0, the resonant frequencies are f<sub>n</sub>=nc/(2$`\sqrt{\epsilon _r}`$L), for n=1,2, …, where $`\epsilon _r`$ is the dielectric constant of the transmission line; for our system a resonance occurs every 125 MHz. For small C<sub>x</sub>, the n-th resonant frequency changes by: $`\mathrm{\Delta }f_nf_n\frac{cC_xZ_0}{L}`$. Since C<sub>x</sub> depends on the distance between the inner conductor and the sample, this equation implies that $`\mathrm{\Delta }`$f can be used to determine the topography of the sample. When the sample is very close to the probe, one expects that C<sub>x</sub> $``$ A$`\epsilon _0`$/h, so that $`\mathrm{\Delta }ff_nc\epsilon _0AZ_0/hL`$, where $`A`$ is the area of the center conductor of the probe. On the other hand, when the sample is far from the probe, one must resort to numerical simulation to find C<sub>x</sub>(h). To find C<sub>x</sub>(h) for the coaxial probe geometry, we solved Poisson’s equation using a finite difference method on a 200 by 150 cell grid. The overall experimental behavior is qualitatively well described by the numerical simulation, including the weak frequency shifts observed at large separation. For convenience, we parameterize our measured calibration curves with an empirical function to easily transform measured frequency shifts to absolute heights. The topographic imaging capabilities of the system were demonstrated by imaging an uneven, metallic sample: a U.S. quarter-dollar coin. First, the frequency shift was recorded as a function of position over the entire sample. We next measured the frequency shift versus height at a fixed position above a flat part of the sample - see Fig. 2 of reference \- and determined the transfer function h($`\mathrm{\Delta }`$f). Using h($`\mathrm{\Delta }`$f), we then transformed the frequency-shift image into a topographic surface plot - see Fig. 3(b) of reference . As a result of this work, we can quantitatively account for the contribution of topography to frequency-shift images. Our technique allows for a height resolution of 55 nm for a 30-$`\mu `$m probe-sample separation and about 40 $`\mu `$m at a separation of 1.75 mm . The technique is simple and should be readily extendible to non-metallic samples, smaller probes and closer or farther separations. ### 5.2 Sheet resistance Based on simple reasoning, we expect that the losses of the sample will affect the quality factor Q of the microscope. A simple situation to study is that of a resistive thin-film on a dielectric substrate with a thickness much less than the skin depth. In this case one can show that the sample presents a sheet resistance R<sub>x</sub> = $`\rho `$/t to the microscope, where $`\rho `$ is the thin-film resistivity and t is the local thickness of the film. To determine the relationship between Q<sub>0</sub> and sample sheet resistance ($`R_X`$), we used a variable-thickness aluminum thin-film on a glass substrate . The cross-section of the thin-film is wedge-shaped, implying a spatially-varying sheet resistance. Using a probe with a 500-$`\mu `$m-diameter center conductor in non-contact mode, and selecting a resonance of the microscope with a frequency of 7.5 GHz, we acquired frequency-shift and Q<sub>0</sub> data. We then cut the sample into narrow strips to take two-point resistance measurements and determine the local sheet resistance. We find that Q<sub>0</sub> reaches a maximum as $`R_X0`$; as $`R_X`$ increases, Q<sub>0</sub> drops due to loss from currents induced in the sample, reaching a minimum around $`R_X=660\mathrm{\Omega }/\mathrm{}`$ for a height of 50 $`\mu `$m. Similarly, as $`R_X\mathrm{}`$, Q<sub>0</sub> increases due to diminishing currents in the sample. This means that $`R_X`$ is a double-valued function of Q<sub>0</sub>. This presents a problem for converting the measured Q<sub>0</sub> to $`R_X`$. However, $`R_X`$ is a single-valued function of the frequency shift , allowing one to use the frequency-shift data to determine which branch of the $`R_X(`$Q$`)`$ curve should be used. To explore the capabilities of our system, we scanned a thin-film of YBa<sub>2</sub>Cu<sub>3</sub>O<sub>7-δ</sub> (YBCO) on a 5-cm-diameter sapphire substrate at room temperature . The film was deposited using pulsed laser deposition with the sample temperature controlled by radiant heating. The sample was rotated about its center during deposition, with the $``$3-cm-diameter plume held at a position halfway between the center and the edge. The thickness of the YBCO thin-film varied from about 100 nm at the edge to 200 nm near the center. Figure 6 shows two microwave images of the YBCO sample. The frequency shift \[Fig. 6(a)\] and Q<sub>0</sub> \[Fig. 6(b)\] were acquired simultaneously, using a probe with a 500-$`\mu `$m-diameter center conductor at a height of approximately 50 $`\mu `$m above the sample. The scan took approximately 10 minutes to complete, with raster lines 0.5 mm apart. The frequency shifts in Fig. 6(a) are relative to the resonant frequency of 7.5 GHz when the probe was far away ($`>`$1 mm) from the sample; the resonant frequency shifted downward by more than 2.2 MHz when the probe was above the center of the sample. Noting that the resonant frequency drops monotonically between the edge and the center of the film, and that the resonant frequency is a monotonically increasing function of sheet resistance , we conclude that the sheet resistance decreases monotonically between the edge and the center. The frequency shift and Q<sub>0</sub> images \[Fig. 6(a) and (b)\] differ slightly in the shape of the contour lines. This is due to the 300-$`\mu `$m-thick substrate being warped, causing a variation of a few microns in the probe-sample separation during the scan. In practice, the frequency shift ($`\mathrm{\Delta }`$f) is dominated by topography, but also has contributions from R<sub>x</sub>. The change in Q is dominated by sheet resistance, but also has contributions from the topography. To deal with this, we have devised a way to deconvolve these influences and produce images of sheet resistance which are not contaminated by topographic features, and topographic images which are not contaminated by sheet resistance variations. The results are shown in Fig. 7. Figure 7(b) confirms that the film does, indeed, have a lower sheet resistance near the center, as was intended when the film was deposited. We note that the sheet resistance does not have a simple radial dependence, due to either non-stoichiometry or defects in the film. Figure 7(a) shows that the wafer is, indeed, warped, with a higher ridge running from upper right to lower left. After scanning the YBCO film, we patterned it and made four-point dc resistance measurements throughout the wafer. The dc sheet resistance had a spatial dependence identical to the microwave data in Fig. 7(b) . However, the absolute values were approximately twice as large as the microwave results, most likely due to degradation of the film during patterning. We also found from surface profilometer traces that the topography image of Fig. 7(a) is quantitatively correct. We have estimated the sheet resistance sensitivity for the microscope as $`\mathrm{\Delta }R_X/R_X`$ = $`6.4\times 10^3`$, for $`R_X=100\mathrm{\Omega }/\mathrm{}`$ using a probe with a 500-$`\mu `$m-diameter center conductor at a height of 50 $`\mu `$m and a frequency of 7.5 GHz . The sensitivity scales with the capacitance between the probe center conductor and the sample ($`C_X`$); increasing the diameter of the probe center conductor and/or decreasing the probe-sample separation would improve the sensitivity. ### 5.3 Dielectric Constant The ability to image variations in relative permittivity or dielectric constant $`\epsilon _r`$ is useful for both fundamental and applied reasons. For example, the dielectric properties of thin-film ferroelectric materials are of interest in studying the finite-size effect on the ferroelectric phase transition. In thin-film microelectronics, testing for variations in dielectric constant can be used for quality control or to develop better growth techniques. Also, knowledge of the dielectric constant at microwave frequencies is of great importance for the design of broadband circuits and future generations of high-speed microprocessors. #### Linear dielectric response - Non-Contact Mode We began our studies of dielectric materials with measurements in non-contact mode with a blunt probe (Fig. 3(a)) . To calibrate the system for dielectric measurements, we constructed a test sample by placing six pieces of different dielectric material into the bottom of a square plastic mold and pouring epoxy into the mold. In addition, silicone adhesive was used to hold down each piece. After the epoxy cured, the test sample was removed from the mold, polished and positioned on the XY table. The materials embedded in the epoxy were silicon, glass microscope slide, SrTiO<sub>3</sub>, Teflon, sapphire and LaAlO<sub>3</sub>. All six pieces were approximately 500-$`\mu `$m thick and about 6 mm x 8 mm in size. The overall thickness of the test sample was 6 mm. We measured the frequency shift $`\mathrm{\Delta }`$f versus height h above the six pieces, which have dielectric constants ranging from 2.1 to about 230. We also tested the epoxy which has an unknown dielectric constant. Each piece, as well as the probe, was flat and smooth on the scale of 5 $`\mu `$m, as judged by an optical microscope. For these measurements, we used a probe with a 480-$`\mu `$m center-conductor diameter and a source frequency of 9.08 GHz. For each scan, the probe was first brought in contact with a dielectric and the frequency shift $`\mathrm{\Delta }`$f was recorded as the height was systematically increased. Samples with the largest dielectric constant produced the largest frequency shift, as expected . The largest shift we observed was $``$26.2 MHz, when the probe was in contact with a SrTiO<sub>3</sub> sample with $`\epsilon _r`$ $``$ 230 . The smallest shift we found was -1.2 MHz when the probe was in contact with a Teflon sample with $`\epsilon _r`$ $``$ 2.1 . The frequency shift is essentially zero above 1 mm and saturates when the probe-sample distance is smaller than a few microns. We used the above information to construct an empirical calibration curve that directly relates the frequency shift to the dielectric constant $`\epsilon _r`$. In order to construct the calibration curve we took the difference between the frequency shift at two different heights, h<sub>1</sub> and h<sub>2</sub>, i.e., f<sub>d</sub>=$`\mathrm{\Delta }`$f(h<sub>2</sub>)-$`\mathrm{\Delta }`$f(h<sub>1</sub>), where h<sub>2</sub> is far away (h<sub>2</sub> $`>`$ 1000 $`\mu `$m). By taking the difference, we eliminated the effect of drift in the microwave source frequency. Proceeding this way for the test samples, we constructed two calibration curves of f<sub>d</sub> versus $`\epsilon _r`$ \- see Fig. 2 of reference \- one curve for h<sub>1</sub>=10 $`\mu `$m and h<sub>2</sub>=1.1 mm, and the other for h<sub>1</sub>=100 $`\mu `$m and h<sub>2</sub>=1.1 mm. We then parameterized each calibration curve with an empirical function, allowing us to easily transform any measured frequency shift to a dielectric constant. From these curves we found that we can enhance the sensitivity to the dielectric constant considerably by using a small probe height. On the other hand, at closer probe-sample separations the influence of topographic features will be enhanced. Hence, it is very important to either control the height of the probe or to de-convolve the topography from the resulting frequency shift and Q images taken in non-contact mode. #### Linear dielectric response - Contact Mode To avoid the issue of topographic features corrupting the dielectric imaging, we have developed a contact-mode version of dielectric imaging . We use the near-field scanning microwave microscope with a sharp protruding center conductor. The probe tip, which has a radius of curvature $`1\mu `$m - see inset to Fig. 2 and Fig. 3(b) - is held fixed, while the sample is supported by a spring-loaded cantilever applying a controlled normal force of about 50 $`\mu `$N between the probe tip and the sample . Due to the concentration of the microwave fields at the tip, the boundary condition of the resonator, and hence, the resonant frequency $`f_0`$ and quality factor Q are perturbed depending on the dielectric properties of the region of the sample immediately beneath the probe tip. We have shown that the spatial resolution of the microscope in this mode of operation is about 1 $`\mu `$m . To observe the microscope’s response to sample dielectric permittivity $`ϵ_r`$, we monitored the frequency-shift signal while scanning samples with known $`ϵ_r`$. With well-characterized 500-$`\mu `$m-thick bulk dielectrics, we observed that the microscope frequency shift monotonically increases in the negative direction with increasing sample permittivity, as noted above . We observed similar behavior with thin-film dielectric samples. For comparison, we did a finite-element calculation of the rf electric field near the probe tip. Because the probe-tip radius is much less than the wavelength $`(\lambda 4`$ cm at 7 GHz), a static calculation of the electric field is sufficient . Cylindrical symmetry further simplifies the problem to two dimensions. We represent the probe tip as a cone with a blunt end, held at a potential of $`\varphi =1`$ V. Using relaxation methods we solved Poisson’s equation for the potential, $`^2\varphi =0`$, on a rectangular grid representing the region around the probe tip. The two variables we used to represent the properties of the probe were the aspect ratio of the probe tip ($`\alpha dz/dr`$) and the radius $`r_0`$ of the blunt end. Since the sample represents a small perturbation to the resonator, we can use perturbation theory to find the change in the resonant frequency: $$\frac{\mathrm{\Delta }f}{f}\frac{ϵ_0}{4W}_{V_S}(ϵ_{r2}ϵ_{r1})𝐄_1𝐄_2𝑑V,$$ (2) where $`𝐄`$<sub>1</sub> and $`𝐄`$<sub>2</sub>, and $`ϵ_{r1}`$ and $`ϵ_{r2}`$ are the unperturbed and perturbed electric fields, and relative permittivities of two samples, respectively; $`W`$ is the energy stored in the resonator, and the integral is over the volume of the sample. We compute an approximate $`W`$ using the fact that the loaded quality factor of the resonator is $`Q_L=\omega _0W/P_{loss}`$, where $`\omega _0`$ is the resonant frequency, and $`P_{loss}`$ is the power loss in the resonator. Using four bulk samples with known relative dielectric permittivities between 2.1 and 305, and fixing $`r_0=(0.6\mu `$m$`)/\alpha `$, we used $`\alpha `$ as a fitting parameter to obtain agreement between the model results from Eq. (2) and our data at 7.2 GHz; we found agreement to within 10 % for several different probe tips, with $`1.0<\alpha <1.7`$. To extend this calibration model to thin films, we extend the finite-element calculation to include a thin-film on top of the dielectric sample substrate. Once a probe’s $`\alpha `$ parameter is determined using the bulk calibration described above, we use the thin-film model combined with Eq. (2), integrating over the volume of the thin-film, to obtain a functional relationship between $`\mathrm{\Delta }f`$ and the dielectric permittivity of the thin-film. Using the thin-film model, we found that for high-permittivity ($`ϵ_r>50`$) thin films, the microwave microscope is primarily sensitive to the in-plane component of the permittivity tensor. Figure 8(a) shows a quantitative permittivity $`ϵ_r`$ image of a sample consisting of a 440-nm SrTiO<sub>3</sub> (STO) thin-film on a 500-$`\mu `$m LaAlO<sub>3</sub> (LAO) substrate. The film was made by pulsed laser deposition at 700 C, in 200 mTorr of O<sub>2</sub>. The film is paraelectric at room temperature. In the microwave (7.2 GHz) permittivity image \[Fig. 8(a)\], the film shows a dielectric constant on the order of 180 over most of its area (20 $`\mu `$m by 20 $`\mu `$m here). Several low-permittivity defects on the film also are visible. From atomic force microscopy measurements we find that these features are second-phase laser particles deposited on top of the STO film. Their greater thickness is part of the reason the relative permittivity is lower in the image. Figure 8(b) shows a quantitative permittivity $`ϵ_r`$ image of a sample consisting of a 275-nm Ba<sub>0.60</sub>Sr<sub>0.40</sub>TiO<sub>3</sub> (BST) thin-film on a 500-$`\mu `$m LAO substrate. The film also was made by pulsed laser deposition at 700 C, in 200 mTorr of O<sub>2</sub>. The film is paraelectric at room temperature. It shows a larger average dielectric constant than STO, but also contains laser particles. The fine structure in Figs. 8 may be indicative of varying dielectric properties from grain to grain in the films. #### Nonlinear dielectric response in contact mode In order to measure the local dielectric tunability of thin films, we apply a dc electric field to the sample by voltage biasing ($`V_{bias}`$) the probe tip - see Fig. 2. A grounded metallic counterelectrode layer immediately beneath the dielectric thin-film acts as a ground plane. To prevent the counterelectrode from dominating the microwave measurement, thus minimizing its effect on the microwave fields - we ignored the counterelectrode in our static field model - the sheet resistance of the counterelectrode should be as high as possible. In our case, we use low carrier density La<sub>0.95</sub>Sr<sub>0.05</sub>CoO<sub>3</sub> with a thickness of 100 nm, giving a sheet resistance $`400\mathrm{\Omega }/\mathrm{}`$. We have confirmed by experiment and model calculation that the contribution of the counterelectrode to the frequency shift is small ($`\mathrm{\Delta }f<30`$ kHz) relative to the contribution from a dielectric thin-film with thickness $`>`$ 100 nm ($`\mathrm{\Delta }f>200`$ kHz). We can examine the tunability of the dielectric properties at a fixed point on the sample and generate a hysteresis loop . As expected, the permittivity goes down and the Q goes up when a voltage is applied. To image the tunability of the sample, we change the bias voltage applied to the probe tip and image the dielectric constant again. A tunability figure of merit can be defined with this data. In general, one wants a highly tunable material which is not very lossy. Hence, the figure of merit to maximize is K = ($`\mathrm{\Delta }\epsilon _r`$/$`\epsilon _r`$)/tan$`\delta `$, where $`\mathrm{\Delta }\epsilon _r`$ is the change in dielectric constant upon tuning with some fixed electric field, and tan$`\delta `$ is the dielectric loss at zero bias. In our case we can define a similar figure of merit with the raw data as K = ($`\mathrm{\Delta }f`$/$`f`$)$`Q`$, where $`\mathrm{\Delta }f`$ is the frequency shift due to tuning a 400-nm-thick Ba<sub>0.60</sub>Sr<sub>0.40</sub>TiO<sub>3</sub> (BST) thin-film with a bias of 2.5 V, and Q is the unbiased quality factor of the microscope. Figure 9 shows the resulting figure of merit image over a 20-$`\mu `$m by 13-$`\mu `$m area. The main region of very low tunability (white area) is most likely a laser-deposited particle of non-paraelectric material. However, there are smaller regions of low tunability which are separated by just a few microns from regions of very good tunability. This demonstrates the power of the scanning microwave microscope to delve into the microstructure-property relationship on the microscopic scale. The microwave technique we use is sensitive to both film thickness and permittivity. As a result, the permittivity of the large defects in Fig. 8(a) is underestimated due to the change in film thickness at these locations, which we have confirmed with our model. However, by examining an AFM image, topographic features can be readily distinguished from permittivity features. We have further shown that in most of the images the topography is too small to account for a significant contribution to the observed permittivity contrast. We can calculate the sensitivity of the microwave microscope by observing the noise in the dielectric permittivity and tunability data. For a 370-nm-thick film on a 500-$`\mu `$m-thick LAO substrate, with an averaging time of 40 ms, we find that the relative dielectric permittivity sensitivity is $`\mathrm{\Delta }ϵ_r=2`$ at $`ϵ_r=500`$, and the tunability sensitivity is $`\mathrm{\Delta }(dϵ_r/dV)=0.03`$ V<sup>-1</sup> ## 6 Future Prospects Having demonstrated quantitative imaging of losses, topography and dielectric constant on the micron length scale and below, what is next for this impressive instrument? Ongoing research is focused on a number of issues. The first is an effort to improve the spatial resolution of the microscope while maintaining its quantitative capabilities. The frontier seems to be improved field-enhancement techniques, similar to those now being pursued in near-field scanning optical microscopy. The exploitation of resonant electromagnetic phenomena combined with cantilever techniques is another tool for improving spatial resolution. Another important issue is the utilization of the microwave microscope’s broad frequency bandwidth. This can be used to examine frequency-dependent phenomena, such as electronic and magnetic dynamics in correlated electron systems. Superconducting materials present an interesting problem for near-field microwave microscopes. Already progress has been made in imaging the transition temperature in superconducting thin films . Although a superconducting microscope is probably required to image surface impedance on the micron scale, it may be possible to learn more from conventional microscopes which study nonlinearities . Our group also has imaged electric fields in the near-field of operating superconducting microwave devices at 77 K and above . The effort is focused on isolating the local sources of nonlinearity under operating conditions. Many other possibilities of electrodynamics experiments remain to be exploited. What is clear is that a new era of local electromagnetic experiments has begun. ## 7 Conclusions We have given a short introduction to the new field of near-field microwave electrodynamics measurements. Five major developments of near-field microwave microscopy have been identified. We have presented a thorough introduction to the form of microscopy we have developed at the University of Maryland. Quantitative images of metallic sheet resistance, topography, dielectric constant and dielectric tunability have been presented. Great potential exists for new quantitative electrodynamics measurements on ever finer length scales. ### 7.1 Acknowledgements This work could not have been possible without the assistance of S. K. Dutta, B. J. Feenstra, W. Hu, A. Schwartz, J. Lee and A. Thanawalla. This work has been supported by the Maryland/NSF Materials Research Science and Engineering Center on Oxide Thin Films DMR 9632521, as well as NSF grant # ECS-9632811, an NSF SBIR (DMI-9710717) subcontract from Neocera, Inc. and by the Maryland Center for Superconductivity Research.
warning/0001/cond-mat0001368.html
ar5iv
text
# Antilocalization in a 2D Electron Gas in a Random Magnetic Field ## Abstract We construct a supersymmetric field theory for the problem of a two-dimensional electron gas in a random, static magnetic field. We find a new term in the free energy, additional to those present in the conventional unitary $`\sigma `$-model, whose presence relies on the long-range nature of the disorder correlations. Under a perturbative renormalization group analysis of the free energy, the new term contributes to the scaling function at one-loop order and leads to antilocalization. A question of long-standing controversy has been the influence of a random, static magnetic field on a two-dimensional electron gas. This problem is relevant to a wide range of physical situations, including the quantum Hall effect in the half-filled Landau level and high-Tc superconductivity in the gauge-field description . Following the application of a succession of analytical methods to this problem (e.g. ), Aronov et al. constructed a $`\sigma `$-model for 2D electron sea in a delta-correlated random magnetic field (RMF). They found that the $`\sigma `$-model was identical to the one derived previously for disordered systems with broken time-reversal invariance (unitary ensemble), and therefore concluded that all states are localized as for the unitary ensemble. In particular, they did not find the additional term in the $`\sigma `$-model proposed in Ref. ; this term appeared in Ref. as a result of an incorrect relation between a correlator of Hall conductivities and the average longitudinal conductivity. Although recently some arguments in favor of delocalization by the RMF were presented , it is apparently commonly believed that all states of a 2D electron gas in the RMF are localized and Ref. gives a proper description of the model. In this paper, we reconsider the problem presenting a more careful derivation of a proper $`\sigma `$-model. It represents a generalization of the supermatrix $`\sigma `$-model to the case of a disorder potential with long-range spatial correlations. For the RMF model, a new term (different from the one of Ref. ) appears in the free energy, additional to those present in the conventional unitary $`\sigma `$-model. The presence of this term is a consequence of the breaking of a “gauge” symmetry due to an anomaly at large momenta. The long-range nature of the vector potential correlations corresponding to the RMF are crucial for our derivation. Under a perturbative renormalization group (RG) analysis of the free energy, the new term contributes to the scaling function at one-loop order and results in the antilocalizing behaviour. A related model which displays electron delocalization is the random flux model which describes, for example, electron hopping on a bipartite lattice structure with link disorder (see e.g. ). Here a tendency towards delocalization is displayed as the band center is approached , due to the existence of a chiral symmetry at the band center. However investigations of a lattice formulation of the RMF problem (equivalent to the random flux model at energies far from the band center) remain controversial: conclusions are divided between the localization of all states and the existence of a critical region . The existence of a metal-insulator transition for an RMF model with a spatially-correlated magnetic field has also been asserted recently . It is also of interest that, even for the case of one dimension, recent analytical and numerical work supports the existence of a metal-insulator transition in the presence of sufficiently long-ranged disorder correlations. It is well-known that spin-orbit impurities in 2D also lead to antilocalization , although by an entirely different mechanism than that considered here. In the following we consider the Hamiltonian, $`=(ie𝐀/c)^2/(2m)ϵ_F.`$ (1) A delta-correlated magnetic field leads to a vector potential, $`𝐀`$, with a long-ranged and transversal correlator, $`A^i(𝐫)A^j(𝐫^{})=2(mc/e)^2V^{ij}(𝐫𝐫^{})`$ where $`V_𝐪^{ij}`$ $`=`$ $`\nu _F^2\gamma {\displaystyle \frac{1}{(q^2+\kappa ^2p_F^2)}}\left(\delta ^{ij}{\displaystyle \frac{q^iq^j}{q^2}}\right),`$ (2) for some strength $`\gamma `$. Here $`ϵ_F`$ and $`v_F`$ are the Fermi-energy and Fermi-velocity and $`\kappa 1`$ is a cutoff that renders finite the otherwise-infinite range of the disorder correlations. At the end of calculations one should take the limit $`\kappa 0`$. This cutoff appears in the divergent single-particle lifetime: for example, the simple Born approximation yields $`\tau _{\mathrm{BA}}^1=2ϵ_F\gamma /\kappa `$ while the self-consistent Born approximation (SCBA) yields the less divergent $`\tau _{\mathrm{SCBA}}^1=2ϵ_F(\gamma \mathrm{ln}(1/\kappa )/\pi )^{1/2}`$. The inverse transport time however remains convergent, $`\tau _{\mathrm{tr}}^1=ϵ_F\gamma `$. We emphasize that our method remains valid for an arbitrary choice of the correlator $`V^{ij}(𝐫)`$, with minor modifications for scalar rather than vector potential disorder. We now derive the free energy functional. Following standard methods and notation , an averaged product of Green’s functions may be expressed as a functional integral weighted by a Lagrangian that is quadratic in an eight-component, supersymmetric $`\psi `$-field, $$=i\overline{\psi }(𝐫)({}_{0}{}^{}\frac{ie}{mc}\tau _3𝐀.\frac{e^2}{mc^2}A^2)\psi (𝐫)d^2𝐫.$$ (3) Here $`{}_{0}{}^{}=ϵ+ϵ_F+^2/(2m)\omega \mathrm{\Lambda }/2`$, and $`\tau _3`$ and $`\mathrm{\Lambda }`$ are the $`z`$-Pauli matrix in time-reversal and retarded-advanced spaces. We average over $`𝐀`$ (neglecting the term in $`A^2`$ as it gives only a slowly varying contribution to the chemical potential) to induce a term in the Lagrangian that is quartic in the $`\psi `$-fields and, in contrast to the case of short-range impurities, non-local in position. Decoupling this term via an 8$`\times `$8 matrix $`Q(𝐫,𝐫^{})`$ and integrating over the $`\psi `$-fields, we come to the Lagrangian $``$ $`=`$ $`{\displaystyle }[{\displaystyle \frac{1}{2}}\mathrm{Str}\mathrm{ln}(i_0+\widehat{V}_{𝐫,𝐫^{}}\stackrel{~}{Q}(𝐫,𝐫^{}))`$ (5) $`+{\displaystyle \frac{1}{4}}\mathrm{Str}\left(Q(𝐫,𝐫^{})\widehat{V}_{𝐫,𝐫^{}}Q(𝐫^{},𝐫)\right)]d^2𝐫d^2𝐫^{},`$ with $`\widehat{V}_{𝐫,𝐫^{}}=(1/2)_{ij}V^{ij}(𝐫𝐫^{})(_𝐫^i_𝐫^{}^i)(_𝐫^j_𝐫^{}^j)`$. Also $`\stackrel{~}{Q}=Q_{}+iQ_{}`$, where $`Q=Q_{}+Q_{}`$ and $`Q_{}`$ ($`Q_{}`$) commutes (anticommutes) with $`\tau _3`$. The $`Q`$-matrix satisfies the standard symmetries $`Q(𝐫,𝐫^{})=\overline{Q}(𝐫^{},𝐫)=KQ^+(𝐫^{},𝐫)K`$, with $`K`$ defined as in Ref. . To search for the saddle-point to Eq. (5), we take $`Q_{}=0`$ and $`Q`$ to depend on $`𝐫𝐫^{}`$ only. After Fourier transforming, the saddle-point equation reads $`Q_𝐩=g_𝐩`$, where $`\left(ϵ\xi _𝐩{\displaystyle \frac{\omega }{2}}\mathrm{\Lambda }\right)g_𝐩2i{\displaystyle \frac{d^2𝐩_1}{(2\pi )^2}V_{𝐩𝐩_1}^{ij}p_1^ip_1^jQ_{𝐩_1}g_𝐩}=i,`$ and $`\xi _𝐩=𝐩^2/(2m)ϵ_F`$. Writing $`Q_𝐩=\mathrm{\Lambda }_𝐩`$, the required form for $`\mathrm{\Lambda }_𝐩`$ is in general complicated. The analysis simplifies however in the limiting cases of $`\kappa (ϵ_F\tau )^1`$ and $`\kappa (ϵ_F\tau )^1`$, where $`\tau `$ is the mean free time that is to be found from the solution. The latter case is relevant here. In both cases we may take $`\mathrm{\Lambda }_𝐩=i/(\xi _𝐩+i\mathrm{\Lambda }/(2\tau ))`$, while keeping $`ϵ_F\tau 1`$. In the limit $`\kappa (ϵ_F\tau )^1`$ one comes to the SCBA solution for the scattering lifetime, $`\tau =\tau _{\mathrm{SCBA}}`$. To consider fluctuations around the saddle-point, we separate them into three types: (a) hard massive, associated with the fluctuations of the eigenvalues of the $`Q`$ and the cooperons, with the mass $`\tau ^1`$ that is very large in the limit $`\kappa 0`$, (b) soft massive, with the mass $`\tau _{\mathrm{tr}}^1`$ that remains finite, and (c) massless. The condition $`ϵ_F\tau 1`$ ensures that fluctuations of type (a) are irrelevant and so $`Q^2(𝐫,𝐫^{})=1`$ and $`[Q,\tau _3]=0`$. To obtain a functional for types (b) and (c), we write $$Q_𝐩(𝐑)=U\left(𝐑\right)Q_𝐩^{\left(0\right)}\overline{U}\left(𝐑\right);Q_𝐩^{\left(0\right)}=V_𝐧(𝐑)\mathrm{\Lambda }_𝐩\overline{V}_𝐧\left(𝐑\right)$$ (6) with $`𝐑=(𝐫+𝐫^{})/2`$ and $`𝐧=𝐩/|p|`$. In Eq. (6), $`U\left(𝐑\right)`$ is independent of $`𝐧`$, and $`Q_𝐩^{\left(0\right)}\left(𝐑\right)`$ contains only non-zero harmonics in $`𝐧`$ around the Fermi surface. The supermatrix $`Q_𝐩^{\left(0\right)}\left(𝐑\right)`$ can be parametrized, for example, as $$Q_𝐩^{\left(0\right)}\left(𝐑\right)=\mathrm{\Lambda }_𝐩\left(1+iP_𝐧\left(𝐑\right)\right)\left(1iP_𝐧\left(𝐑\right)\right)^1$$ (7) with $`𝑑𝐧P_𝐧=0`$, $`\overline{P}_𝐧=P_𝐧`$, $`\{P_𝐧,\mathrm{\Lambda }\}=0`$. Both $`U\left(𝐑\right)`$ and $`P_𝐧\left(𝐑\right)`$ vary slowly with $`𝐑`$, that is, on length scales longer than $`l=v_F\tau `$. The parametrization for $`Q_𝐩\left(𝐑\right)`$ guarantees that the final free energy functional describing fluctuations will be invariant under global rotations $`U\left(𝐑\right)U_0U\left(𝐑\right)`$ with $`\overline{U}_0=U_0^1`$ independent on coordinates, this invariance being present in the initial Lagrangian, Eq. (3). As we are interested in the low-frequency behavior of transport coefficients the massless modes are of greatest interest for us. However, we cannot simply neglect the soft massive modes (b), and our aim now is to integrate them out, thus reducing the free energy to an effective functional containing only $`U\left(𝐑\right)`$. Due to the soft mass of the (b) modes it is sufficient to apply a Gaussian approximation in $`P_𝐧\left(𝐑\right)`$. Higher order terms give a small contribution provided the inequality $`ϵ_F\tau _{\mathrm{tr}}1`$ is fulfilled. The crucial feature of this integration is that, as a consequence of the long-range correlations of the disorder, Eq. (2), it requires a cutoff at large momentum $`k`$ (short distance) which remains in the final functional. There is some freedom in how to choose the cutoff for fluctuations of $`P_𝐧\left(𝐑\right)`$ as the symmetry under the global rotations is already guaranteed by the choice of parametrization and there are no other symmetries to be respected in the initial Lagrangians, Eq. (3, 5). The simplest way is to write this cutoff as the distance $`l_0`$, where $`l_{\mathrm{tr}}=v_F\tau _{\mathrm{tr}}l_0l`$, below which the fluctuations are suppressed $$P_{\mathrm{𝐧𝐤}}=0\text{ for }\left|𝐤\right|>l_0^1$$ (8) where $`P_{\mathrm{𝐧𝐤}}`$ are Fourier components of $`P_𝐧\left(𝐑\right)=_𝐤P_{\mathrm{𝐧𝐤}}\mathrm{exp}(i𝐤.𝐑\tau _3)`$. As we have to calculate only Gaussian integrals over $`P_{\mathrm{𝐧𝐤}}`$, Eq. (8) is very convenient for explicit computation. At the end of the calculation we will take $`l_0l`$. We remark that to recover the conventional $`\sigma `$-model for short-range (delta-correlated) impurities , we may set $`V_𝐧=1`$ in Eq. (6) so that $`Q`$ is a function of $`𝐑`$ only. It then follows that the free energy possesses a local “gauge” invariance (LGI) under the transformation $`U(𝐑)U(𝐑)h(𝐑)`$, where $`[h(𝐑),\mathrm{\Lambda }]=0`$ (note the term “gauge” is used here in a separate sense from that of the original vector potential). It is the loss of this invariance in the case of long-ranged disorder that leads finally to antilocalization. Before explaining this point we present the expression for the free energy functional $`F[U]`$ obtained after integration over the non-zero harmonics $`P_{\mathrm{𝐧𝐤}}`$: $`F[U]`$ $`=`$ $`{\displaystyle }d^2𝐫[{\displaystyle \frac{\pi \nu }{8}}\mathrm{Str}[D(Q\left(𝐫\right))^2+2i\omega \mathrm{\Lambda }Q\left(𝐫\right)]`$ (10) $`{\displaystyle \frac{\beta }{32\pi }}\left[\mathrm{Str}(\tau _3\mathrm{\Lambda }\overline{U}\left(𝐫\right)U\left(𝐫\right))\right]^2],`$ where $`D=v_F^2\tau _{\mathrm{tr}}/2`$ is the classical diffusion coefficient, $`Q\left(𝐫\right)=U\left(𝐫\right)\mathrm{\Lambda }\overline{U}\left(𝐫\right)`$, and $`\beta =\tau _{\mathrm{tr}}/\left(2\tau \right)1`$. The first line in Eq. (10) corresponds to the conventional $`\sigma `$-model for the unitary ensemble that predicts localization. This part has been derived in Ref. for the model in question. The second line in Eq. (10) is the main result of the present paper. The first line respects the LGI but the term in the second line violates this invariance and leads eventually to antilocalization. At first glance, the violation of the LGI seems impossible because, on replacing $`U(𝐑)U(𝐑)h(𝐑)`$ in Eqs. (5, 6), we can change the variables of integration $`P_𝐧\left(𝐑\right)\overline{h}\left(𝐑\right)P_𝐧\left(𝐑\right)h\left(𝐑\right)`$ when integrating over the non-zero harmonics. This would remove $`h\left(𝐑\right)`$ from the final expression for $`F[U]`$. However, this change of the variables would replace Eq. (8) by a more complicated expression for the cutoff. The LGI is therefore guaranteed only if the cutoff is not important, that is, when no ultraviolet divergencies appear in the integral over the non-zero harmonics. Now let us explain the most important steps of the derivation of Eq. (10). First, we perform an expansion of the logarithm in the Lagrangian (5), in both $`\overline{T}_𝐧T_𝐧`$ (where $`T_𝐧=UV_𝐧`$) and the collision integral (deviation of $`\widehat{V}_{𝐫,𝐫^{}}\stackrel{~}{Q}(𝐫,𝐫^{})`$ from its saddle-point value). If we keep to first order in both of these terms, we recover the Muzykantskii and Khmelnitskii $`\sigma `$-model with ballistic disorder . Although such an expansion is sufficient to describe a long-ranged potential (for which $`\tau _{\mathrm{tr}}\tau `$), for the general case including a short-range potential we require the corrections to the free energy in Ref. obtained by continuing this expansion to second-order in both terms. The resulting terms in the free energy then complete the list of those that contribute to Gaussian order in $`P_𝐧`$. We find it convenient to Fourier transform from $`P_{\mathrm{𝐧𝐤}}`$ to angular harmonics, $`P_{\mathrm{𝐧𝐤}}=_mP_{m𝐤}\mathrm{exp}\left(im\varphi \tau _3\right)`$. Expressing the collision integral in terms of these harmonics, we find that, as well as the SCBA scattering lifetime described previously, a whole series of lifetimes associated with successive harmonics appears in the free energy. We define the $`m`$th lifetime $`\tau ^{(m)}`$ by $`{\displaystyle \frac{\mathrm{\Lambda }}{2\tau ^{(m)}}}=2{\displaystyle \frac{d^2𝐩_1}{(2\pi )^2}V_{𝐩𝐩_1}^{ij}p_1^ip_1^j\mathrm{\Lambda }_{p_1}\mathrm{cos}(m\phi )},`$ where $`\phi `$ is the angle between $`𝐩`$ and $`𝐩_1`$, so that $`\tau ^{(0)}`$ coincides with $`\tau `$ and $`1/\tau _{\mathrm{tr}}=1/\tau 1/\tau ^{(1)}`$. The free energy becomes $`F=F_0+F_{}+F_{}+F_{\mathrm{unit}}`$ with $`F_0`$ $`=`$ $`\pi \nu {\displaystyle }{\displaystyle \frac{d^2𝐤}{(2\pi )^2}}{\displaystyle \underset{m=1}{\overset{\mathrm{}}{}}}\mathrm{Str}[P_{m𝐤}P_{m,𝐤}{\displaystyle \frac{\tau ^{(m)}}{\tau }}({\displaystyle \frac{1}{\tau }}{\displaystyle \frac{1}{\tau ^{(m)}}})`$ $`+{\displaystyle \frac{iv_F}{2}}\mathrm{\Lambda }(P_{m𝐤}P_{m1,𝐤}\overline{k}^{}+P_{m+1,𝐤}P_{m,𝐤}\overline{k})],`$ $`F_{}`$ $`=`$ $`\pi \nu v_F{\displaystyle }d^2𝐫{\displaystyle \underset{m=1}{\overset{\mathrm{}}{}}}\mathrm{Str}[\mathrm{\Phi }_x^{}\tau _3\mathrm{\Lambda }(P_mP_{m1}`$ $`+P_{m+1}P_m)i\mathrm{\Phi }_y^{}\mathrm{\Lambda }(P_mP_{m1}P_{m+1}P_m)],`$ $`F_{}`$ $`=`$ $`{\displaystyle \frac{i\pi \nu v_F}{2}}{\displaystyle }d^2𝐫\mathrm{Str}[\mathrm{\Phi }_x^{}\tau _3\mathrm{\Lambda }(P_1+P_1)`$ $`+i\mathrm{\Phi }_y^{}\mathrm{\Lambda }(P_1P_1)],`$ $$F_{\mathrm{unit}}=\frac{\pi \nu }{8}d^2𝐫\mathrm{Str}\left[D_0(Q)^2+2i\omega \mathrm{\Lambda }Q\right],$$ (11) where $`\nu `$ is the density of states, $`\overline{k}=k_x+ik_y\tau _3`$, $`\mathrm{\Phi }\left(𝐫\right)=\overline{U}\left(𝐫\right)U\left(𝐫\right)`$, and $`\mathrm{\Phi }^{}`$ $`(\mathrm{\Phi }^{})`$ are the components of $`\mathrm{\Phi }`$ that commute (anti-commute) with $`\mathrm{\Lambda }`$. Also $`Q=U\left(𝐫\right)\mathrm{\Lambda }\overline{U}\left(𝐫\right)`$ and $`D_0=v_F^2\tau /2`$. As an intermediate step we have rescaled $`P_mP_m\tau ^{(m)}/\tau `$. Having derived the free-energy functional for the fluctuations in $`Q_𝐧\left(𝐫\right)`$, Eq. (11), the next step is to integrate over the non-zero harmonics, represented by $`P_𝐧`$, to derive a functional in terms of only $`U\left(𝐫\right)`$. We may check immediately that for short-range (delta-correlated) disorder, the conventional unitary sigma-model is recovered: in this case, $`\tau ^{(m)}\mathrm{}`$ for $`m0`$, the $`P_m`$ fields become infinitely massive and only the terms in $`F_{\mathrm{unit}}`$ remain as required. For the case of the delta-correlated RMF, we find instead by explicit evaluation that $`1/\tau 1/\tau ^{(m)}=(2m1)/\tau _{\mathrm{tr}}`$, and $`\tau ^{(m)}/\tau 1`$, for $`m\kappa ^1`$. We then average the terms in $`\mathrm{\Phi }`$ with respect to the functional $`F_0`$. The leading contribution, containing no more than two gradient operators, comes from the second-order cumulant of $`(F_{}+F_{})`$, according to $`F_0+F_{}+F_{}(1/2)(F_{}+F_{})^2_{F_0}`$. To perform the averaging we apply what amounts to a straightforward generalization of the standard contraction rules to include the angular harmonic space (see e.g. ). The contribution of the mixed term $`F_{}F_{}`$ vanishes. The terms in $`F_{}`$ lead to a dressing of the bare diffusion coefficient, $`D_0`$, appearing in $`F_{\mathrm{unit}}`$: $`D_0D=v_F^2\tau _{\mathrm{tr}}/2`$. Taking into account only the contributions from $`F_{\mathrm{unit}}`$ and $`F_{}`$ corresponds to the calculation of Ref. and gives the first line of Eq. (10). The terms in $`F_{}`$ induce the new term in the free energy $`F[U]`$ (second line in Eq. (10)), where the dimensionless coefficient $`\beta `$ can be written as $$\beta =2\pi _{\left|𝐬\right|<s_0}\frac{d^2s}{\left(2\pi \right)^2}\left(\frac{^2}{s_x^2}+\frac{^2}{s_y^2}\right)\mathrm{Tr}\mathrm{ln}\widehat{L}\left(\overline{s}\right)$$ (12) Here $`𝐬=𝐤l_{\mathrm{tr}}/2`$ is a rescaled momentum variable which must be cut off at some upper limit set by $`l_0`$ as discussed above: we have $`s_0=l_{\mathrm{tr}}/(2l_0)1`$. The semi-infinite, tridiagonal matrix $`\widehat{L}`$ in Eq. (12) has the entries $`\left(\widehat{L}(\overline{s})\right)_{m,n}=(2m1)\delta _{m,n}+i\overline{s}\delta _{m+1,n}+i\overline{s}^{}\delta _{m,n+1}`$ for $`1m,n\kappa ^1`$ and $`\overline{s}=s_x+is_y`$. We see that the integral in Eq. (12) is determined only by a contribution of the boundary $`\left|𝐬\right|=s_0`$ ($`det\widehat{L}`$ depends on $`s=|𝐬|`$ only): $$\beta =\left[s\frac{d}{ds}\left(\mathrm{ln}det\widehat{L}(\overline{s})\right)\right]_{s=s_0}.$$ (13) In the case where the cutoff $`l_0`$ is unimportant, such as for sufficiently short-range disorder, this term becomes identically zero and the LGI is preserved as required. In the more general case, we see that $`\beta `$ is determined by the asymptotes of the function $`det\widehat{L}\left(\overline{s}\right)`$, that in turn are determined by the range of the disorder correlations. For the RMF model, we have succeeded in calculating the determinant of $`\widehat{L}`$ (for $`\kappa 0)`$ analytically as follows. If we set $`\widehat{L}_N`$ as the truncation of $`\widehat{L}`$ to $`1m,nN`$ and introduce $`M_N(s)=det\widehat{L}_N(s)/((2N1)!!)`$, we find the recurrence relation $`M_N=M_{N1}+{\displaystyle \frac{s^2}{(2N1)(2N3)}}M_{N2},`$ with the boundary conditions $`M_0=M_1=1`$. Next we introduce the function $`f(z,s)=_{N=0}^{\mathrm{}}M_N(s)z^N`$ and derive from the recurrence relation a second-order differential equation for $`f(z,s)`$. As a result we come to the simple formula $`\underset{N\mathrm{}}{lim}M_N(s)=\mathrm{cosh}(s)`$ (14) and hence $`\beta (s_0)=s_0\mathrm{tanh}(s_0)s_01`$, which should be used in Eq. (10). We now perform a perturbative RG analysis on the free energy (10) (see Ref. ): we separate $`U`$ into slow ($`\stackrel{~}{U}`$) and fast ($`U_0`$) components, $`U=\stackrel{~}{U}U_0`$, and expand $`U_0`$ perturbatively in $`P`$, where $`U_0=(1iP)^{1/2}(1+iP)^{1/2}`$. As for the conventional unitary $`\sigma `$-model, the diffusion coefficient $`D`$ receives a negative correction at two-loop order. To derive the correction it is necessary to go below two dimensions and analytically continue the finite result to 2D. For the free energy (10), the diffusion coefficient gains a further, positive correction originating from the new term in $`\beta `$, this time at only one-loop ($`O(P^2)`$) order. At the same time, $`\beta `$ is not renormalized in this order. The scaling function for $`t=8/(\pi \nu D)`$ in 2D becomes $$\frac{dt}{d\mathrm{ln}\lambda }=\frac{1}{2}t^3(\beta 1)+O(t^4),$$ (15) where $`\lambda `$ is the running momentum cutoff for the $`P`$-fields, with the solution $$t(\omega )=t_0[(1+(\beta 1)t_0^2\mathrm{ln}(1/\omega \tau _{tr})]^{1/2}.$$ (16) Finally we take the cutoff $`l_0l`$. If we also cut off the range of the disorder correlations at the system size, $`L`$, we have $`\kappa 1/(p_FL)`$. This gives $`\beta =\tau _{\mathrm{tr}}/2\tau (\mathrm{ln}(p_FL)/\pi \gamma )^{1/2}1`$. We see that the resistivity decreases as the frequency $`\omega `$ is lowered and eventually vanishes, corresponding to antilocalization. In deriving the above scaling function it is crucial that the function $`\beta (s_0)`$ for $`s_0\mathrm{}`$ survives dimensional regularization below 2D. In fact, the exponential dependence of $`det\widehat{L}(s)`$ on $`s`$ for large $`s`$ ensures that $`\beta `$ remains finite for any dimension $`d>1`$. For any finite range disorder correlations the function $`det\widehat{L}\left(s\right)`$ would grow algebraically. Extending Eq. (12, 13) to $`d<2`$, we see that $`\beta `$ vanishes in this case and the conventional unitary $`\sigma `$-model is recovered. In summary, we have considered the problem of a 2D electron gas in a random, static magnetic field and found that the resistivity vanishes in the limit of small frequencies, a behaviour which may be classified as antilocalization. This result may have far-reaching consequences for gauge field models in the theory of strongly correlated systems and lead to other interesting applications. We are grateful to I. Aleiner, A.Altland, V.E. Kravtsov, and A.I. Larkin for many helpful discussions and gratefully acknowledge the financial support of the Sonderforschungsbereich 237.
warning/0001/hep-ph0001209.html
ar5iv
text
# Anomalous 𝑓₁ exchange in vector meson photoproduction asymmetries ## Abstract We perform an analysis of the elastic production of vector mesons with polarized photon beams at high energy in order to investigate the validity of a recently proposed dynamical mechanism based on the dominance of the $`f_1`$ trajectory at large momentum transfer. The density matrix characterizing the angular distributions of the vector meson decays is calculated within an exchange model which includes the Pomeron and the $`f_1`$. The asymmetries of these decays turn out to be very useful to disentangle the role of these exchanges since their effect depends crucially on their quantum numbers which are different. The observables analyzed are accessible with present experimental facilities. preprint: SNUTP-00-001 FTUV-00-0118 hep-ph/0001209 Exclusive photoproduction experiments of vector mesons have become powerful tools for testing diffractive mechanisms at high energy . Regge theory has been successful in describing the diffractive production in terms of the Pomeron exchange mechanism. Donnachie and Landshoff showed that, by assuming the Pomeron-photon analogy and introducing a form factor for the coupling of the Pomeron to quarks, the diffractive vector meson production with real and virtual photons could be described successfully by the soft Pomeron exchange. The soft Pomeron exchange governs this process for small $`|t|`$ and fulfills the $`s`$-channel helicity conservation, a consequence of the old vector dominance model and a requirement of the experimental data . However, at larger $`|t|`$, the soft Pomeron alone cannot explain the recent ZEUS data on elastic vector meson photo- and electroproduction and new contributions seem necessary. For example Donnachie and Landshoff introduce in addition to the soft Pomeron the hard Pomeron with the trajectory $`\alpha _P^{}=1.44+0.1t`$ describing in this way the data up to $`|t|2`$ GeV<sup>2</sup>. Recently we have suggested a new anomalous Regge trajectory with high intercept $`\alpha _{f_1}(0)1`$ and small slope $`\alpha _{f_1}^{}0`$ .It is not well known how the Pomeron arises from QCD, although it seems quite plausible that it is related to the conformal anomaly of the theory . In the same way we do not yet know how the anomalous $`f_1`$ trajectory arises from QCD, but we have strong suspicions that the origin of its physical relevance lies in its relation to the axial anomaly of the theory. This trajectory has the quantum numbers $`P=C=+1`$ and the signature $`\sigma =1`$ while the Pomeron carries $`P=C=\sigma =+1`$. In Ref. we have shown that the $`f_1`$ exchange describes the vector meson photoproduction data at large energy and momentum transfer.<sup>\**</sup><sup>\**</sup>\**The $`f_1`$ trajectory also gives natural explanation to the behavior of the cross sections of elastic hadron-hadron scattering at large $`|t|`$ and furthermore its contribution to the flavor singlet part of the spin-dependent structure function $`g_1`$ at low $`x`$ region gives a new explanation to the proton spin problem. In this model, the soft Pomeron is dominant at $`|t|1`$ GeV<sup>2</sup> while the $`f_1`$ exchange dominates the large $`|t|`$ region, $`|t|1`$ GeV<sup>2</sup>. In order to understand the details of the mechanisms involved in our model it is important to investigate other physical quantities which can distinguish between the two exchanges, i.e., the Pomeron and the $`f_1`$, in a clear way. Diffractive production of vector mesons by polarized photon beams seems to be the appropriate tool for such purpose as we will show hereafter. One of the important features of the new anomalous $`f_1`$ trajectory is its odd signature, which should discriminate it from the Pomeron which has even signature. Therefore the contribution from this new exchange can be disentangled from the Pomeron contribution in spin-dependent processes. In order to investigate the very specific features of the $`f_1`$ trajectory contribution we consider vector meson production with polarized photon beams and its decay into pseudoscalar mesons. We find that the asymmetries of the vector meson decays described by the soft Pomeron and $`f_1`$ exchanges are drastically different from the predictions obtained with the soft and hard Pomeron exchanges. Our starting point is the density matrix of the vector meson production by photons from proton targets, $$\rho _{\lambda _V\lambda _V^{}}=\frac{1}{N}\underset{\lambda _N^{},\lambda _\gamma ,\lambda _N,\lambda _\gamma ^{}}{}T_{\lambda _V\lambda _N^{},\lambda _\gamma \lambda _N}\rho (\gamma )_{\lambda _\gamma \lambda _\gamma ^{}}T_{\lambda _V^{}\lambda _N^{},\lambda _\gamma ^{}\lambda _N}^{},$$ (1) where $`T`$ is the $`T`$-matrix element of elastic vector meson photoproduction process, $`\lambda `$’s are the polarization states of the particles, and $`N`$ is the normalization factor defined as $$N=\frac{1}{2}\underset{\lambda \text{’s}}{}|T_{\lambda _V\lambda _N^{},\lambda _\gamma \lambda _N}|^2.$$ (2) The photon density matrix $`\rho (\gamma )`$ is given by $$\rho (\gamma )=\frac{1}{2}(1+𝐏_\gamma 𝝈),$$ (3) where $`𝝈`$ is the Pauli matrix and $`𝐏_\gamma `$ specifies the polarization of linearly polarized photons and is given by $$𝐏_\gamma =p_\gamma (\mathrm{cos}2\mathrm{\Phi },\mathrm{sin}2\mathrm{\Phi },0),$$ (4) where $`\mathrm{\Phi }`$ denotes the angle between the photon polarization vector and the vector meson production plane, and $`p_\gamma `$ denotes the magnitude of the polarization ($`0p_\gamma 1`$). The decay angular distribution of the vector meson in its rest frame reads $`{\displaystyle \frac{d𝒩}{d\mathrm{cos}\vartheta d\phi }}W(\mathrm{cos}\vartheta ,\phi ,\mathrm{\Phi })=W^0(\mathrm{cos}\vartheta ,\phi )+{\displaystyle \underset{i=1}{\overset{3}{}}}P_\gamma ^i(\mathrm{\Phi })W^i(\mathrm{cos}\vartheta ,\phi ),`$ (5) where $`\vartheta `$ and $`\phi `$ are the polar and azimuthal angles of the direction of flight of one pseudoscalar meson in the vector meson rest frame. As in the literature, we use the Gottfried-Jackson frame as the vector meson rest frame, where the $`z`$ axis is in the direction of the incident photon as seen in this frame. (See Refs. for details.) The explicit forms of $`W^\alpha `$ are $`W^0(\mathrm{cos}\vartheta ,\phi )`$ $`=`$ $`{\displaystyle \frac{3}{4\pi }}\{{\displaystyle \frac{1}{2}}(1\rho _{00}^0)+{\displaystyle \frac{1}{2}}(3\rho _{00}^01)\mathrm{cos}^2\vartheta \sqrt{2}\text{Re}\rho _{10}^0\mathrm{sin}2\vartheta \mathrm{cos}\phi `$ (7) $`\text{Re}\rho _{11}^0\mathrm{sin}^2\vartheta \mathrm{cos}2\phi \},`$ $`W^1(\mathrm{cos}\vartheta ,\phi )`$ $`=`$ $`{\displaystyle \frac{3}{4\pi }}\left\{\rho _{11}^1\mathrm{sin}^2\vartheta +\rho _{00}^1\mathrm{cos}^2\vartheta \sqrt{2}\text{Re}\rho _{10}^1\mathrm{sin}2\vartheta \mathrm{cos}\phi \text{Re}\rho _{11}^1\mathrm{sin}^2\vartheta \mathrm{cos}2\phi \right\},`$ (8) $`W^2(\mathrm{cos}\vartheta ,\phi )`$ $`=`$ $`{\displaystyle \frac{3}{4\pi }}\left\{\sqrt{2}\text{Im}\rho _{10}^2\mathrm{sin}2\vartheta \mathrm{sin}\phi +\text{Im}\rho _{11}^2\mathrm{sin}^2\vartheta \mathrm{sin}2\phi \right\},`$ (9) $`W^3(\mathrm{cos}\vartheta ,\phi )`$ $`=`$ $`{\displaystyle \frac{3}{4\pi }}\left\{\sqrt{2}\text{Im}\rho _{10}^3\mathrm{sin}2\vartheta \mathrm{sin}\phi +\text{Im}\rho _{11}^3\mathrm{sin}^2\vartheta \mathrm{sin}2\phi \right\},`$ (10) where $`\rho _{ij}^\alpha `$ are the matrix elements of $`\rho ^\alpha `$ which are defined by $$\rho ^0=\frac{1}{2N}TT^{},\rho ^i=\frac{1}{2N}T\sigma ^iT^{},$$ (11) with $`i=1,2,3`$, whose normalization is $`\text{Tr}\rho ^0=1`$. There are various decay angular distribution functions arising from different photon polarizations, whose measurements determine the vector meson density matrix elements. Interesting quantities in connection with the nature of the exchanged particles are the asymmetries. Depending on the direction of the polarization vector of the linearly polarized photon beams, we define the asymmetry $`\mathrm{\Sigma }`$ as $$\mathrm{\Sigma }\frac{\sigma _{}\sigma _{}}{\sigma _{}+\sigma _{}}=\frac{1}{p_\gamma }\frac{W^L(0,\frac{\pi }{2},\frac{\pi }{2})W^L(0,\frac{\pi }{2},0)}{W^L(0,\frac{\pi }{2},\frac{\pi }{2})+W^L(0,\frac{\pi }{2},0)},$$ (12) where $`\sigma _{}`$ ($`\sigma _{}`$) is the cross section for the symmetric decay of particle pairs produced parallel (normal) to the plane of polarization of the photon. $`W^L`$ represents angular distribution for the decay (5) with linearly polarized photon beams. In terms of the density matrix, it can be written as $$\mathrm{\Sigma }=\frac{\rho _{11}^1+\rho _{11}^1}{\rho _{11}^0+\rho _{11}^0}.$$ (13) Another relevant quantity is the parity asymmetry $`P_\sigma `$, which is defined from the observation that, if either natural ($`P=\sigma `$) or unnatural parity ($`P=\sigma `$) exchange in the $`t`$-channel contributes, one has an additional symmetry , $$T_{\lambda _V\lambda _N^{},\lambda _\gamma \lambda _N}=\pm (1)^{\lambda _V\lambda _\gamma }T_{\lambda _V\lambda _N^{},\lambda _\gamma \lambda _N},$$ (14) from which we get $$\rho _{\lambda \lambda ^{}}^{0(N/U)}=\frac{1}{2}\left[\rho _{\lambda \lambda ^{}}^0(1)^\lambda \rho _{\lambda \lambda ^{}}^1\right].$$ (15) This allows us to define the parity asymmetry by means of $`\sigma ^N`$ and $`\sigma ^U`$, which are the contributions of natural and unnatural parity exchanges to the cross section respectively as $$P_\sigma \frac{\sigma ^N\sigma ^U}{\sigma ^N+\sigma ^U}=2\rho _{11}^1\rho _{00}^1.$$ (16) Therefore when we have only the natural parity exchange we get $`P_\sigma =+1`$, while we obtain $`P_\sigma =1`$ when only the unnatural parity exchange contributes. We apply the above formalism to $`\rho `$ and $`\varphi `$ meson photoproduction with polarized photon beams. We denote the four-momenta of the initial proton by $`p`$, that of the final proton by $`p^{}`$, the photon beam four-momentum by $`q`$, and that of the produced vector meson by $`q_V`$. The matrix element for the soft Pomeron exchange part reads $`T_{\lambda _Vm^{},\lambda _\gamma m}^P`$ $`=`$ $`i12\sqrt{4\pi \alpha _{\mathrm{em}}}\beta _uG_P(w^2,t)F_1(t){\displaystyle \frac{m_V^2\beta _f}{f_V}}{\displaystyle \frac{1}{m_V^2t}}\left({\displaystyle \frac{2\mu _0^2}{2\mu _0^2+m_V^2t}}\right)`$ (18) $`\times \left\{\overline{u}_m^{}(p^{})\overline{)}qu_m(p)\epsilon _V^{}(\lambda _V)\epsilon _\gamma (\lambda _\gamma )\left[q\epsilon _V^{}(\lambda _V)\right]\overline{u}_m^{}(p^{})\gamma _\mu u_m(p)\epsilon _\gamma ^\mu (\lambda _\gamma )\right\},`$ (19) where the vector meson and the photon helicities are denoted by $`\lambda _V`$ and $`\lambda _\gamma `$ while $`m`$ and $`m^{}`$ are the spin projections of the initial and final nucleon, respectively. The remaining quantities are defined by $`G_P(w^2,t)`$ $`=`$ $`\left({\displaystyle \frac{w^2}{s_0}}\right)^{\alpha _P(t)1}\mathrm{exp}\left\{{\displaystyle \frac{i\pi }{2}}[\alpha _P(t)1]\right\},`$ (20) $`F_1(t)`$ $`=`$ $`{\displaystyle \frac{4m_p^22.8t}{(4m_p^2t)(1t/0.71)^2}},`$ (21) with $`w^2=(2W^2+2m_p^2m_V^2)/4`$ and $`W^2=(p+q)^2`$. $`m_p`$ represents the proton mass, while $`m_V`$ the vector meson masses, and the Pomeron trajectory is $`\alpha _P(t)=1.08+\alpha _P^{}t`$ with $`\alpha _P^{}=1/s_0=0.25`$ GeV<sup>-2</sup>. We use $`\mu _0^2=1.1`$ GeV<sup>2</sup>, $`\beta _u=\beta _d=2.07`$ GeV<sup>-1</sup> and $`\beta _s=1.45`$ GeV<sup>-1</sup>. The vector meson decay constant is represented by $`f_V`$. The $`f_1`$ exchange amplitude reads $`T_{\lambda _Vm^{},\lambda _\gamma m}^{f_1}`$ $`=`$ $`ig_{f_1V\gamma }g_{f_1NN}F_{f_1NN}(t)F_{f_1V\gamma }(t){\displaystyle \frac{m_V^2}{tm_{f_1}^2}}ϵ_{\mu \nu \alpha \beta }q^\mu \epsilon _V^\nu (\lambda _V)\epsilon _\gamma ^\alpha (\lambda _\gamma )`$ (23) $`\times \left(g^{\beta \delta }{\displaystyle \frac{(pp^{})^\beta (pp^{})^\delta }{m_{f_1}^2}}\right)\overline{u}_m^{}(p^{})\gamma _\delta \gamma _5u_m(p),`$ where the $`f_1V\gamma `$ coupling constants are determined from the $`f_1`$ decay: $`g_{f_1\rho ^0\gamma }=0.94`$ GeV<sup>-2</sup> and $`g_{f_1\varphi \gamma }=0.18`$ GeV<sup>-2</sup>. The form factors are $`F_{f_1NN}(t)=1/(1t/m_{f_1}^2)^2`$ with $`m_{f_1}`$ ($`=1.285`$ GeV) defining the $`f_1`$ mass and $`F_{f_1V\gamma }(t)=(\mathrm{\Lambda }_V^2m_{f_1}^2)/(\mathrm{\Lambda }_V^2t)`$ with $`\mathrm{\Lambda }_\rho =1.5`$ GeV and $`\mathrm{\Lambda }_\varphi =1.8`$ GeV. We refer for the details of the amplitudes to Ref. . In Fig. 1 we show the differential cross section for $`\rho `$ photoproduction at $`\gamma p`$ c.m. energy $`W=94`$ GeV, which is the kinematical region of the ZEUS experiments. The different role of the Pomeron and $`f_1`$ exchanges is apparent: The Pomeron dominates at small $`|t|`$ while the $`f_1`$ gives the major contribution at larger $`|t|`$. The differential cross section for $`\varphi `$ photoproduction can be found in Ref. . Figures 2 and 3 show the density matrices defined in Eq. (11) for $`\rho `$ and $`\varphi `$ photoproduction for the same energy. The figures emphasize the diverse features of the Pomeron and $`f_1`$ exchanges arising as a consequence from their different symmetry properties (14). The inclusion of the $`f_1`$ exchange changes the signs of some density matrix elements at large $`|t|`$ where the $`f_1`$ exchange dominates the process. This feature is responsible for the dramatic difference in the asymmetries between the two approaches. We give predictions for the parity asymmetry $`P_\sigma `$ in Fig. 4. We obtain identical result for the $`\mathrm{\Sigma }`$ asymmetry of Eq. (12). $`\mathrm{\Sigma }`$ is not related unambiguously to natural and unnatural parity exchange, but it becomes equivalent to $`P_\sigma `$ if the helicity-flip amplitudes are suppressed as in our case. Furthermore, $`P_\sigma =\pm 1`$ implies $`\mathrm{\Sigma }=\pm 1`$ although the reverse implication is not always true . Because of its natural parity, the Pomeron exchange leads to $`P_\sigma =+1`$ while the $`f_1`$ exchange gives $`P_\sigma =1`$ due to its unnatural parity. Therefore in Fig. 4 one can view the relative strength of the two exchanges as a function of $`|t|`$. In $`\rho `$ photoproduction the two exchanges are comparable in magnitude at $`|t|1`$ GeV<sup>2</sup>, which leads to the vanishing of $`P_\sigma `$ in this region. Below this region, the Pomeron dominates and the asymmetry approaches $`+1`$, while it becomes $`1`$ for $`|t|>2`$ GeV<sup>2</sup> where the $`f_1`$ dominance is clearly established. Although, as shown above, the best way to distinguish the two mechanisms in vector meson production is to use the polarized photon beams, similar information can be obtained from vector meson electroproduction with unpolarized electron beam experiments at small $`Q^2`$, which can be performed at present electron facilities. Data on the density matrix in vector meson electroproduction by fixed-target experiments is available . Recently the H1 and ZEUS Collaborations reported data on the density matrix elements in $`\rho ^0`$ electroproduction at higher energy. Both seem to be consistent with the Pomeron exchange model. However it should be noted that these data were obtained only in the region of small $`|t|`$, say $`|t|0.6`$ GeV<sup>2</sup>, with large errors. In this region the natural parity exchange (Pomeron exchange) dominates and the $`f_1`$ exchange contribution is small, so it is not possible to draw any definite conclusion on the effect of the $`f_1`$ exchange from these limited data set. Since the $`f_1`$ exchange alters the predictions of the Pomeron exchange at large $`|t|`$, it is necessary to measure the $`|t|`$-dependence of the density matrices up to $`|t|2`$ GeV<sup>2</sup>, and this may clarify the nature of the exchanged trajectory which is responsible for vector meson production at large $`|t|`$. In summary, we have shown that the new anomalous unnatural-parity $`f_1`$ exchange leads to significant $`|t|`$ dependence of the $`P_\sigma `$ and $`\mathrm{\Sigma }`$ asymmetries in polarized vector meson photoproduction. The recent claim of Donnachie and Landshoff that the relatively large $`|t|`$ data of the ZEUS experiments could be explained by including the hard Pomeron will lead to a very different prediction on these asymmetries and can be discriminated from the $`f_1`$ exchange process. We have good reason to believe that the existence of the anomalous $`f_1`$ exchange in vector meson production is deeply related to the properties of the axial anomaly in QCD . Therefore the investigation of the decay asymmetries in vector meson production by polarized photon or (un)polarized lepton beams at present experimental facilities such as CERN, DESY and Fermilab will shed light on our understanding of the diffractive processes from the fundamental structure of QCD. We are grateful to J. A. Crittenden, S. B. Gerasimov, and A. I. Titov for fruitful discussions. Y.O. and D.P.M. were supported in part by the KOSEF through the CTP of Seoul National University. V.V. was supported by DGICYT-PB97-1227. N.I.K. thanks the Department of Physics of Seoul National University for the warm hospitality.
warning/0001/astro-ph0001494.html
ar5iv
text
# Electromagnetic Energy for a Charged Kerr Black Hole in a Uniform Magnetic Field ## Abstract With the Komar mass formula we calculate the electromagnetic energy for a charged Kerr black hole in a uniform magnetic field. We find that the total electromagnetic energy takes the minimum when the Kerr black hole possesses a non-zero net charge $`Q=2\xi B_0J_H`$ where $`B_0`$ is the strength of the magnetic field, $`J_H`$ is the angular momentum of the black hole, $`\xi `$ is a dimensionless parameter determined by the spin of the black hole. Whether an astrophysical black hole can possess a net electric charge is an interesting and important question. For an astrophysical black hole without magnetic field the answer seems to be clear: usually the black hole cannot possess much net electric charges since otherwise the black hole will selectively accrete particles with opposite sign of charges from ambient material and be quickly neutralized . (However, if accretion onto a black hole produces a luminosity close to Eddington limit, the black hole can acquire a net charge due to the different radiation drag on electrons and ions .) But for a black hole immersed in a magnetic field — which is believed to occur in many astrophysical systems — the situation is different and the answer is quite unclear. Wald has shown that when a Kerr black hole is immersed in a uniform magnetic field aligned with its rotation axis, the hole acquires a net electric charge $`Q_W=2B_0J_H`$, where $`B_0`$ is the strength of the magnetic field, $`J_H`$ is the angular momentum of the black hole (throughout the paper we use the geometric units $`G=c=1`$) . Wald derived his result from the requirement that the “injection energy” along the symmetry axis should be zero for the equilibrium state. However, it can be shown that the “injection energy” defined by Wald in depends on the path along which the injection is made and the effect of off-axis accretion of charges is unclear. So it remains a question what is the equilibrium state for a Kerr black hole in a uniform magnetic field and whether the equilibrium state acquires a net charge. Ruffini and Treves have analyzed the problem of a magnetized rotating sphere in flat spacetime . They have found that the sphere acquires a non-zero net charge in order to minimize the total electromagnetic energy of the system. This is similar to the conclusion of Wald but the acquired charge has opposite sign. The method of extremizing the total energy is better than that of “injection energy” since the former is a global approach which doesn’t depend on the details of injection path and accretion process. Therefore in this paper we try to calculate the total electromagnetic energy for a charged Kerr black hole in a uniform magnetic field. Energy (mass) is well-defined for a stationary system which is asymptotically flat and has an asymptotic timelike Killing vector . However, if the uniform magnetic field extends to infinity in space, the total electromagnetic energy diverges. To obtain a finite electromagnetic energy we truncate the electromagnetic field with a spherical surface. The black hole sits at the center of the sphere, the radius of the sphere is much larger than the radius of the black hole horizon. Inside the sphere the magnetic field is uniform but outside the sphere the magnetic field decreases quickly with increasing radius. (Currents are induced on the surface of the sphere to connect the magnetic fields inside and outside the sphere.) We will calculate the total electromagnetic energy inside the sphere using the Komar mass formula and show that when the total electromagnetic energy takes the minimum the black hole acquires a non-zero net charge $`Q=2\xi B_0J_H=\xi Q_W`$, where the dimensionless parameter $`\xi `$ is a function of $`a/M_H`$ and $`0\xi \left[\frac{3}{2}\left(2+\pi \right)\right]^10.13`$ for $`0a/M_H1`$, $`M_H`$ is the mass of the black hole and $`a=J_H/M_H`$. (The value of $`\xi `$ is independent of the radius of the truncation sphere in the limit that the radius of the sphere is much larger than the radius of the black hole.) So the Wald state (which corresponds to $`\xi =1`$) doesn’t minimize the total electromagnetic energy. Indeed, the energy of the Wald state is higher than that of the state with no charge. For a charged rotating black hole immersed in a uniform magnetic field with the electromagnetic field being sufficiently weak ($`Q^2/M_H^21`$ and $`B_0^2M_H^21`$), the spacetime can be described with Kerr metric. (In other words, the electromagnetic field is treated as test field in the background of Kerr spacetime.) In Boyer-Lindquist coordinates, the Kerr metric is $`ds^2=\left(1{\displaystyle \frac{2M_Hr}{\mathrm{\Sigma }}}\right)dt^2{\displaystyle \frac{4M_Har}{\mathrm{\Sigma }}}\mathrm{sin}^2\theta dtd\varphi +{\displaystyle \frac{\mathrm{\Sigma }}{\mathrm{\Delta }}}dr^2+\mathrm{\Sigma }d\theta ^2+{\displaystyle \frac{C\mathrm{sin}^2\theta }{\mathrm{\Sigma }}}d\varphi ^2,`$ (1) where $`M_H`$ is the Komar mass of the black hole, $`a`$ is the specific angular momentum of the black hole (the angular momentum of the black hole is $`J_H=M_Ha`$), and $`\mathrm{\Delta }=r^22M_Hr+a^2,\mathrm{\Sigma }=r^2+a^2\mathrm{cos}^2\theta ,C=(r^2+a^2)^2\mathrm{\Delta }a^2\mathrm{sin}^2\theta .`$ (2) The electromagnetic vector potential is $`A^a={\displaystyle \frac{B_0}{2}}\left[\left({\displaystyle \frac{}{\varphi }}\right)^a+2a\left({\displaystyle \frac{}{t}}\right)^a\right]{\displaystyle \frac{Q}{2M_H}}\left({\displaystyle \frac{}{t}}\right)^a,`$ (3) where $`B_0`$ is the strength of the external magnetic field and $`Q`$ is the electric charge of the black hole. The electromagnetic field $`F_{ab}`$ is given by $`F_{ab}=_aA_b_bA_a.`$ (4) The tensor indexes are raised and lowered with the Kerr metric $`g_{ab}`$. Since $`\left(\frac{}{t}\right)^a`$ and $`\left(\frac{}{\varphi }\right)^a`$ are the Killing vectors of Kerr spacetime and the Ricci tensor $`R_{ab}=0`$ for Kerr metric, $`F_{ab}`$ solves vacuum Maxwell’s equations $`_aF^{ab}=0`$ . The stress-energy tensor of the electromagnetic field is $`T_{ab}={\displaystyle \frac{1}{4\pi }}\left(F_{ac}F_b^c{\displaystyle \frac{1}{4}}g_{ab}F_{de}F^{de}\right).`$ (5) The trace of the stress-energy tensor of the electromagnetic field is $`T=g_{ab}T^{ab}=0`$. With the Komar mass formula , the total mass (energy) for the Kerr black hole and the electromagnetic field (with suitable truncation as described earlier) is $`M=2{\displaystyle _\mathrm{\Sigma }}\left(T_{ab}{\displaystyle \frac{1}{2}}Tg_{ab}\right)n^a\left({\displaystyle \frac{}{t}}\right)^b𝑑V+{\displaystyle \frac{1}{4\pi }}\kappa 𝒜+2\mathrm{\Omega }_HJ_H,`$ (6) where $`n^a={\displaystyle \frac{1}{\alpha }}\left[\left({\displaystyle \frac{}{t}}\right)^a+{\displaystyle \frac{2aM_Hr}{C}}\left({\displaystyle \frac{}{\varphi }}\right)^a\right],\alpha \left({\displaystyle \frac{\mathrm{\Delta }\mathrm{\Sigma }}{C}}\right)^{1/2},`$ (7) is the unit future-pointing normal to the hypersurface $`\mathrm{\Sigma }(t=\mathrm{constant})`$, $`dV`$ is the volume element on $`\mathrm{\Sigma }`$, $`𝒜=4\pi \left(r_H^2+a^2\right)`$ is the area of the horizon, $`r_H=M_H+\sqrt{M_H^2a^2}`$ is the radius of the horizon, $`\kappa =\left(r_HM_H\right)/\left(2M_Hr_H\right)`$ is the surface gravity of the black hole, and $`\mathrm{\Omega }_H=a/\left(2M_Hr_H\right)`$ is the angular velocity of the black hole. So the total energy of the electromagnetic field within the truncation sphere is $`_{EM}={\displaystyle _{r_H}^R}𝑑r{\displaystyle _0^\pi }𝑑\theta {\displaystyle _0^{2\pi }}𝑑\varphi ϵ\sqrt{h},ϵ2\left(T_{ab}{\displaystyle \frac{1}{2}}Tg_{ab}\right)n^a\left({\displaystyle \frac{}{t}}\right)^b,`$ (8) where $`\sqrt{h}=\left({\displaystyle \frac{\mathrm{\Sigma }C}{\mathrm{\Delta }}}\right)^{1/2}\mathrm{sin}\theta `$ (9) is the measure of volume on $`\mathrm{\Sigma }`$. The integration over $`r`$ is truncated at $`r=R`$ (the radius of the truncation sphere), otherwise the integration diverges since the magnetic field is asymptotically uniform. Suppose $`R`$ is $`r_H`$ but $`\left(M_H/B_0^2\right)^{1/3}`$ \[so that $`_{EM}M_H`$ and the background spacetime can be well approximated with the Kerr geometry\]. The energy so defined is conserved since the spacetime is stationary. Inserting Eqs. (3 \- 5) into Eq. (8), we obtain $`ϵ={\displaystyle \frac{\alpha }{4\pi \mathrm{\Sigma }^3}}\left(f_0B_0^2+f_1Q^2+f_2B_0Q\right),`$ (10) where $`f_0`$ $`=`$ $`r^5(r2M_H\mathrm{sin}^2\theta )+a^2r^2[3r^2\mathrm{cos}^2\theta +2M_Hr(1+\mathrm{cos}^2\theta )(13\mathrm{cos}^2\theta )`$ (13) $`M_H^2(1+\mathrm{cos}^2\theta )(35\mathrm{cos}^2\theta )]+a^4[3r^2\mathrm{cos}^4\theta 2M_Hr\mathrm{cos}^2\theta (3+\mathrm{cos}^2\theta )+`$ $`M_H^2(1+\mathrm{cos}^2\theta )^2(2\mathrm{cos}^2\theta )]+a^6\mathrm{cos}^6\theta ,`$ $`f_1=r^2+a^2\left(2\mathrm{cos}^2\theta \right),`$ (14) $`f_2`$ $`=`$ $`2a\{r^2(rM_H)(13\mathrm{cos}^2\theta )a^2[r\mathrm{cos}^2\theta (3\mathrm{cos}^2\theta )`$ (16) $`M_H(1+\mathrm{cos}^2\theta )(2\mathrm{cos}^2\theta )]\},`$ and $`_{EM}=_0+\left({\displaystyle \frac{Q^2}{M_H}}\right)F_1+\left({\displaystyle \frac{B_0J_HQ}{M_H}}\right)F_2,`$ (17) where (in the limit $`Rr_H`$) $`_0={\displaystyle \frac{1}{3}}B_0^2R^3+𝒪\left(R^2\right),`$ (18) $`F_1=F_1(s)={\displaystyle \frac{1}{2s^3}}\left(1\sqrt{1s^2}\right)\left[s+2\mathrm{arctan}\left({\displaystyle \frac{s}{1+\sqrt{1s^2}}}\right)\right]+𝒪\left({\displaystyle \frac{1}{R}}\right),`$ (19) $`F_2=F_2(s)={\displaystyle \frac{2}{3s^3}}\left[s(3\sqrt{1s^2}s^2)6(1s^2)\mathrm{arctan}\left({\displaystyle \frac{s}{1+\sqrt{1s^2}}}\right)\right]+𝒪\left({\displaystyle \frac{1}{R^2}}\right),`$ (20) where $`sa/M_H`$ is the spin parameter of the black hole. Since $`0s1`$, $`F_1>0`$ and $`F_20`$ always. Since the total electromagnetic field is the superposition of the external magnetic field $`B_0`$ and the electromagnetic field generated by the charge $`Q`$, the total electromagnetic energy is composed of three parts: (1) the “bare” energy $`_0`$, i.e. the energy of the external magnetic field, which is proportional to $`B_0^2`$; (2) the energy of the electromagnetic field generated by the charge $`Q`$, which is proportional to $`Q^2`$ (c.f. for the electromagnetic energy of a Kerr-Newman geometry); (3) the energy arising from the interaction between the external magnetic field $`B_0`$ and the electromagnetic field of the charge $`Q`$, which is proportional to $`B_0Q`$. Though as $`R\mathrm{}`$ the “bare” energy $`_0`$ diverges, $`F_1`$ and $`F_2`$ converge. So $`F_1`$ and $`F_2`$ don’t depend on where the truncation is if $`Rr_H`$. In particular, the difference in the electromagnetic energy between a charged Kerr black hole in a uniform magnetic field and its uncharged state, $`\mathrm{\Delta }=_{EM}_0`$, is independent of the truncation and so is well-defined. $`_{EM}`$ has a minimum since $`F_1(s)>0`$ for $`0s1`$. By $`_{EM}/Q=0`$ we obtain $`Q=2\xi (s)B_0J_H,`$ (21) where $`\xi (s)={\displaystyle \frac{F_2}{4F_1}}={\displaystyle \frac{s^33s\sqrt{1s^2}+6(1s^2)\mathrm{arctan}\left(\frac{s}{1+\sqrt{1s^2}}\right)}{3\left(1\sqrt{1s^2}\right)\left[s+2\mathrm{arctan}\left(\frac{s}{1+\sqrt{1s^2}}\right)\right]}}.`$ (22) For $`0s1`$ we have $`0\xi \left[\frac{3}{2}\left(2+\pi \right)\right]^10.13`$. $`\xi (s)`$ is plotted in Fig. 1. $`\xi `$ decreases quickly as $`s`$ decreases. As examples: $`\xi (0)=0`$, $`\xi (0.1)3.4\times 10^4`$, $`\xi (0.5)9.9\times 10^3`$, $`\xi (0.9)5.6\times 10^2`$, $`\xi (0.99)0.10`$, and $`\xi (1)0.13`$. So, a Kerr black hole immersed in a uniform magnetic field acquires a non-zero net charge \[given by Eq. (21)\] to attain the minimum electromagnetic energy. Relative to the bare state (i.e. the state with $`Q=0`$), the value of the minimum electromagnetic energy is $`\mathrm{\Delta }_{min}=\left({\displaystyle \frac{F_2^2}{4F_1}}\right){\displaystyle \frac{B_0^2J_H^2}{M_H}}.`$ (23) For a Kerr black hole with $`a/M_H=0.99`$ we have $`Q0.20B_0J_H`$ and $`\mathrm{\Delta }_{min}0.045B_0^2J_H^2/M_H`$. Wald’s result $`Q_W=2B_0J_H`$ does not correspond to the lowest energy state of the electromagnetic field. In fact, the difference in the electromagnetic energy between the Wald state and the uncharged state is $`\mathrm{\Delta }_W=2\left(2F_1+F_2\right){\displaystyle \frac{B_0^2J_H^2}{M_H}},`$ (24) which is always positive for $`0a/M_H1`$. (For a Kerr black hole with $`a/M_H=0.99`$ we have $`\mathrm{\Delta }_W=3.4B_0^2J_H^2/M_H`$.) So the electromagnetic energy for the Wald state is even higher than that of the uncharged state. To see if the charge given by Eq. (21) is important in practice, let’s compare the contribution of the charge $`Q`$ and the magnetic field $`B_0`$ to the magnetic flux through the northern hemi-sphere of the black hole horizon. The magnetic flux contributed by $`Q`$ is $`\mathrm{\Phi }_Q=2\pi aQ/r_H`$. The magnetic flux contributed by $`B_0`$ is $`\mathrm{\Phi }_{B_0}=2\pi B_0M_Hr_H\left(1a^2/r_H^2\right)`$. (The total magnetic flux is $`\mathrm{\Phi }=\mathrm{\Phi }_Q+\mathrm{\Phi }_{B_0}`$.) For $`Q`$ given by Eq. (21) we have $`{\displaystyle \frac{\mathrm{\Phi }_Q}{\mathrm{\Phi }_{B_0}}}={\displaystyle \frac{\xi s^2}{1s^2+\sqrt{1s^2}}},`$ (25) which increases with increasing $`s`$. We find that $`0\mathrm{\Phi }_Q/\mathrm{\Phi }_{B_0}<1`$ if $`0s<0.995`$, $`1<\mathrm{\Phi }_Q/\mathrm{\Phi }_{B_0}<\mathrm{}`$ if $`0.995<s1`$. So the charge given by Eq. (21) is important only if $`a/M>0.995`$. In conclusions, we have calculated the electromagnetic energy for a charged Kerr black hole immersed in a uniform magnetic field. We have found that a non-zero net charge is acquired by the Kerr black hole for attaining the minimum electromagnetic energy. The Wald state isn’t the state with minimum electromagnetic energy. Indeed the electromagnetic energy for the Wald state is even higher than that for the uncharged state. Though the realistic case for an astrophysical black hole is much more complicated than the simple model investigated here due to the appearance of many charged particles in the neighborhood of the black hole, our results have shown that a Kerr black hole in the lowest energy state in an external magnetic field acquires a non-zero net charge. ###### Acknowledgements. I am very grateful to Robert M. Wald for many helpful and stimulating discussions. This work was supported by the NSF grant AST-9819787.
warning/0001/quant-ph0001073.html
ar5iv
text
# Entangled SU(2) and SU(1,1) coherent states ## I Introduction Qubits are the basic elements of quantum information technology, in the same way that bits are basic units of information in computers. Whereas bits are binary digits, qubits are spin-$`1/2`$, or two-level, quantum systems. The advantage of quantum computing over classical computing is the capacity for producing entangled qubits: the large state space available for entangled qubits enables certain problems, thought not to be computable on classical computers, to be solved on quantum computers. A bit can be in an off, or ‘0’, state or an on, or ‘1’, state, but the qubit can be in a superposition of an off, or ‘$`|0`$’, state and an on, or ‘$`|1`$’, state. We can represent such a state (a general qubit) by $`|\theta ,\varphi `$ $`=`$ $`\mathrm{exp}\left[{\displaystyle \frac{\theta }{2}}(\sigma _+e^{i\varphi }\sigma _{}e^{i\varphi })\right]|1`$ (1) $`=`$ $`\mathrm{cos}{\displaystyle \frac{\theta }{2}}|1+e^{i\varphi }\mathrm{sin}{\displaystyle \frac{\theta }{2}}|0,`$ (2) up to a global phase. Here $`\sigma _\pm =\sigma _x\pm i\sigma _y`$ for $`\sigma _x`$ and $`\sigma _y`$ are Pauli matrices. In fact such a state is an SU(2) coherent state, also known as an atomic coherent state . Two qubits, prepared in a product state, could then be expressed as $`|\theta _1,\varphi _1|\theta _2,\varphi _2`$. However, quantum computation is based on the exploitation of entanglement, and such product states are of limited value. The simplest extension of this arbitrary two-qubit product state to a two-qubit entangled state is the unnormalized state $`\mathrm{cos}\theta |\theta _1,\varphi _1|\theta _2,\varphi _2`$ (3) $`+e^{i\varphi }\mathrm{sin}\theta |\theta _1^{},\varphi _1^{}|\theta _2^{},\varphi _2^{},`$ (4) which is a product state for $`\theta `$ a multiple of $`\pi /2.`$ Of course the Bell states $`|\mathrm{\Phi }_\pm `$ $`=`$ $`{\displaystyle \frac{1}{\sqrt{2}}}\left(|0|0\pm |1|1\right),`$ (5) $`|\mathrm{\Psi }_\pm `$ $`=`$ $`{\displaystyle \frac{1}{\sqrt{2}}}\left(|0|1\pm |1|0\right)`$ (6) are special cases of the general state (1.2). The SU(2) coherent states form an overcomplete basis for the Hilbert space, and the qubit states correspond to spin-$`1/2`$ representations of SU(2). There are therefore subtleties concerning this entanglement of non-orthogonal states: such subtleties have been considered with respect to entangled coherent states (or superpositions of multimode coherent states) where the coherent states have been harmonic oscillator coherent states. Our objective here is to introduce and analyze entangled SU(2) and SU(1,1) coherent states. The entangled SU(1,1) coherent states are closely related to the SU(2) coherent states because the algebra su(1,1) and su(2) are so similar. However, there are two commonly considered coherent states for SU(1,1). One SU(1,1) coherent state is the analog of the harmonic oscillator coherent state achieved by displacing the vacuum state and the SU(2) coherent state obtained by “rotating” the lowest- or highest-weight state. The analogous SU(1,1) coherent state is obtained via an SU(1,1) transformation of lowest-weight state. This SU(1,1) coherent state is a member of Perelomov’s category of generalized coherent state, and we refer to this state as a Perelomov SU(1,1) coherent state. The second SU(1,1) coherent state, introduced by Barut and Girardello, is the analog of the harmonic oscillator coherent state being an eigenstate of the annihilation operator; an SU(2) coherent state of this type does not exist due to the SU(2) Hilbert space being finite. We treat entangled SU(1,1) coherent states of both the Perelomov and Barut-Girardello types. As a special case of the entangled Perelomov SU(1,1) coherent state, we obtain superpositions of squeezed vacuum states and entangled squeezed states. Squeezed states are significant in quantum limited measurements, quantum communications and exotic spectroscopy of atoms. In association with the parity operator, this paper first develops parity coherent states and entangled SU(2) and SU(1,1) coherent states in Section II. Section III generalizes the parity coherent states and considers nonlinear SU(2) and SU(1,1) coherent states. In Section IV, we discuss how to represent entangled SU(2) and SU(1,1) coherent states in Fock space, and we have obtained the entangled binomial states, entangled negative binomial states and entangled squeezed states as well as the contraction of entangled SU(2) and SU(1,1) coherent states to entangled harmonic oscillator coherent states. Section V investigates how to generate the entangled coherent states in Hamiltonian systems. Section VI discusses the degree of entanglement for these entangled coherent states, and a conclusion is given in Section VII. The Appendix gives the most general entangled SU(2) and SU(1,1) coherent states with certain special cases, including the quantum Fourier transform state of a product multiparticle SU(2) coherent state acquired by Shor’s algorithm. ## II Entangled SU(2) and SU(1,1) coherent states ### A Entangled coherent states of the harmonic oscillator For the harmonic oscillator, a general unnormalized two-particle entangled coherent state can be expressed as $`\mathrm{cos}\theta |\alpha _1|\alpha _2+e^{i\varphi }\mathrm{sin}\theta |\alpha _1^{}|\alpha _2^{}`$ (7) $`=[\mathrm{cos}\theta \mathrm{exp}(\stackrel{}{\alpha }\stackrel{}{a}^{}\stackrel{}{\alpha }^{}\stackrel{}{a})`$ (8) $`+e^{i\varphi }\mathrm{sin}\theta \mathrm{exp}(\stackrel{}{\alpha }^{}\stackrel{}{a}^{}\stackrel{}{\alpha }^{}\stackrel{}{a})]|0,0,`$ (9) for $`\stackrel{}{a}(a_1,a_2)`$, $`\stackrel{}{\alpha }(\alpha _1,\alpha _2)`$ and $`|0,0=|0_1|0_2`$ the vacuum state. It is particularly helpful to concentrate on the balanced entangled coherent state $$2^{1/2}\left[|\alpha _1|\alpha _2+e^{i\varphi }|\alpha _1|\alpha _2\right],$$ (10) where the word “balanced” refers to each component in the superposition having coefficients of the same magnitude. For $`\varphi =\pi /2`$ such a state can, in principle, be generated via a nonlinear interferometer and is an eigenstate of the canonically transformed pair annihilation operator $$\stackrel{}{\stackrel{~}{a}}\mathrm{\Pi }\stackrel{}{a}$$ (11) with $$\mathrm{\Pi }=e^{i\pi \stackrel{}{a}^{}\stackrel{}{a}},\text{ }\mathrm{\Pi }\stackrel{}{a}\mathrm{\Pi }=\stackrel{}{a}.$$ (12) Thus, the entangled coherent state (2.1) is actually a pair coherent state with respect to the canonically transformed vector annihilation operator, although this pair coherent state does not have the restriction that the number difference of the two modes is fixed. The most general superposition of harmonic oscillator coherent states is $$\frac{d^{2N}\stackrel{}{\alpha }}{\pi ^N}f(\stackrel{}{\alpha })|\stackrel{}{\alpha }$$ (13) with $`|\stackrel{}{\alpha }=|\alpha _1|\alpha _2\mathrm{}|\alpha _N`$ the $`N`$particle coherent state. The superposition (2.5) is entangled when it cannot be expressed as a product state in any representation. We can introduce an entangled SU(2) coherent state as a generalization of (1.2). However, we make this analysis more general than necessary for studying qubits. We wish to treat general irreducible representations $`j`$, where $`j`$ is the angular momentum parameter and can be integer or half-odd integer. The single qubit case corresponds to $`j=1/2`$. A pair of qubits can of course be treated as a single system: in this case we have one $`j=0`$ state (the singlet state) and $`j=1`$ states (the triplet states). The four states together constitute the Bell states (1.3). ### B Entangled SU(2) coherent states The SU(2) coherent state can be expressed as $`|j\text{ }\gamma `$ $``$ $`R(\gamma )|jj`$ (14) $`=`$ $`\mathrm{exp}\left[{\displaystyle \frac{1}{2}}\theta \left(J_+e^{i\varphi }J_{}e^{i\varphi }\right)\right]|jj`$ (15) $`=`$ $`\left(1+|\gamma |^2\right)^j{\displaystyle \underset{m=0}{\overset{2j}{}}}\left({\displaystyle \genfrac{}{}{0pt}{}{2j}{m}}\right)^{1/2}\gamma ^m|jjm,`$ (16) where $`\gamma =\mathrm{exp}(i\varphi )\mathrm{tan}(\theta /2),`$ $`R(\gamma )`$ is the rotation operator, and $`J_{}`$ and $`J_+`$ are lowering and raising operators of the su(2) Lie algebra, respectively. The su(2) generators $`J_\pm `$ and $`J_z`$ satisfy the su(2) commutation relations $$[J_+,J_{}]=2J_z,\text{ }[J_z,J_\pm ]=\pm J_\pm .$$ (17) We can define an su(2) parity operator as $$\mathrm{\Pi }=(1)^{},\text{ }\mathrm{\Pi }^2=1,\text{ }\mathrm{\Pi }^{}=\mathrm{\Pi }.$$ (18) Here $`=J_z+j`$ is the ‘number’ operator such that $$|j,m=(j+m)|j,m,\text{ }0j+m2j.$$ (19) It is easy to see that $$\mathrm{\Pi }J_\pm \mathrm{\Pi }=J_\pm ,\text{ }\mathrm{\Pi }J_z\mathrm{\Pi }=J_z.$$ (20) Using the above equation we have a new su(2) representation: $$[\stackrel{~}{J}_+,\stackrel{~}{J}_{}]=2\stackrel{~}{J}_z,\text{ }[\stackrel{~}{J}_z,\stackrel{~}{J}_\pm ]=\pm \stackrel{~}{J}_\pm ,$$ (21) where $$\stackrel{~}{J}_+=J_+\mathrm{\Pi },\text{ }\stackrel{~}{J}_{}=\mathrm{\Pi }J_{}\text{}\stackrel{~}{J}_zJ_z.$$ (22) We define a new SU(2) coherent state associated with the su(2) algebra (2.11) as $`|j\text{ }\gamma _\mathrm{\Pi }`$ $``$ $`\stackrel{~}{R}(\gamma )|jj`$ (23) $`=`$ $`\mathrm{exp}[{\displaystyle \frac{\theta }{2}}(\stackrel{~}{J}_+e^{i\varphi }\stackrel{~}{J}_{}e^{i\varphi })]|jj.`$ (24) We call $`|j`$ $`\gamma _\mathrm{\Pi }`$ the parity SU(2) coherent state, because the parity operator plays a central role in its definition. This term follows from that of the parity (harmonic oscillator) coherent state. The state $`|j`$ $`\gamma _\mathrm{\Pi }`$ is different from the SU(2) coherent state due to the nontrivial introduction of the su(2) parity operator $`\mathrm{\Pi }`$. The antinormally ordered rotation operator is $$\stackrel{~}{R}(\gamma )=\mathrm{exp}(\gamma \stackrel{~}{J}_{})(1+|\gamma |^2)^{\stackrel{~}{J}_z}\mathrm{exp}(\gamma ^{}\stackrel{~}{J}_+).$$ (25) Using the above equation, we obtain $`|j\text{ }\gamma _\mathrm{\Pi }={\displaystyle \frac{1}{\sqrt{2}}}[e^{i\frac{\pi }{4}}|j\text{ }i(1)^{2j}\gamma `$ (26) $`+e^{i\frac{\pi }{4}}|j\text{ }i(1)^{2j}\gamma ].`$ (27) The SU(2) parity coherent state is a superposition of two SU(2) coherent states with a phase difference $`\pi `$. It is similar to the parity harmonic oscillator coherent states. The general entangled SU(2) coherent state, analogous to the entangled coherent state (2.5), is discussed in Appendix A. Here we introduce a specific SU(2) coherent state by employing the su(2) parity operator (2.8). Let us consider two independent su(2) Lie algebras. From these two algebras, we can define two new algebras $`[\stackrel{~}{J}_+^n,\stackrel{~}{J}_{}^l]`$ $`=`$ $`2\delta _{nl}\stackrel{~}{J}_z^n,\text{ }[\stackrel{~}{J}_z^n,\stackrel{~}{J}_\pm ^l]=\pm \delta _{nl}\stackrel{~}{J}_\pm ^n,`$ (28) $`[\stackrel{~}{J}_{}^n,\stackrel{~}{J}_{}^l]`$ $`=`$ $`0,`$ (29) where $`\stackrel{~}{J}_+^n`$ $`=`$ $`J_+^n\mathrm{\Pi },\text{ }\stackrel{~}{J}_{}^n=\mathrm{\Pi }J_{}^n\text{}\stackrel{~}{J}_z^nJ_z^n,`$ (30) $`\mathrm{\Pi }`$ $`=`$ $`(1)^{\text{ }_1+_2},\text{ }n,l=1,2.`$ (31) It is easy to see that the parity operator $`\mathrm{\Pi }`$ satisfies $`\mathrm{\Pi }^2=1`$ and $`\mathrm{\Pi }^{}=\mathrm{\Pi }`$. These two new su(2) representation are dependent on each other. The SU(2) coherent state of the two original su(2) algebra is $$|j\text{ }\stackrel{}{\gamma }=R(\stackrel{}{\gamma })|jj_1|jj_2,$$ (32) where $`\stackrel{}{\gamma }(\gamma _1,\gamma _2),`$ $`|j`$ $`\stackrel{}{\gamma }|j`$ $`\gamma _1_1|j`$ $`\gamma _2_2,`$ and $`R(\stackrel{}{\gamma })R_1(\gamma _1)R_2(\gamma _2).`$ It is important that both Hilbert spaces concerned in the entanglement are restricted to the same irreducible representation $`j`$ for the entangled coherent states as we introduce them to be well-defined for the $`j=1/2`$ case. We consider entangled qubit states; hence the restriction to the same $`j`$ is of particular value as well. However the SU(2) coherent state for the two new su(2) representations is obtained as $`|j\text{ }\stackrel{}{\gamma }_\mathrm{\Pi }`$ $`=`$ $`\stackrel{~}{R}(\stackrel{}{\gamma })|jj_1|jj_2`$ (33) $`=`$ $`{\displaystyle \frac{1}{\sqrt{2}}}\left(e^{i\frac{\pi }{4}}|j\text{ }i\stackrel{}{\gamma }+e^{i\frac{\pi }{4}}|j\text{ }i\stackrel{}{\gamma }\right)`$ (34) The state $`|j`$ $`\stackrel{}{\gamma }_\mathrm{\Pi }`$ is an entangled SU(2) coherent state with a two-particle parity symmetry. ### C Entangled SU(1,1) coherent states The generators of su(1,1) Lie algebras, $`K_\pm `$ and $`K_z`$, satisfy the commutation relations $$[K_+,K_{}]=2K_z,\text{ }[K_z,K_\pm ]=\pm K_\pm .$$ (35) By analogy to the su(2) case, we can define the su(1,1) parity operator as $$\mathrm{\Pi }=(1)^𝒩,\text{ }\mathrm{\Pi }^2=1,\text{ }\mathrm{\Pi }^{}=\mathrm{\Pi }.$$ (36) where the ‘number’ operator $`𝒩`$ is given by $$𝒩=K_zk,\text{ }𝒩|k\text{ }n=n|k\text{ }n.$$ (37) Here $`|k`$ $`n`$ $`(n=0,1,2,\mathrm{})`$ is the complete orthonormal basis and $`k\{1/2,1,3/2,2,\mathrm{}\}`$ is the Bargmann index labeling the irreducible representation\[$`k(k1)`$ is the value of the Casimir operator\]. Using the su(1,1) parity operator we can introduce a new su(1,1) algebra with generators $$\stackrel{~}{K}_+=K_+\mathrm{\Pi },\text{ }\stackrel{~}{K}_{}=\mathrm{\Pi }K_{}\text{}\stackrel{~}{K}_zK_z.$$ (38) There are two distinct SU(1,1) coherent states to consider. #### 1 Entangled Perelomov SU(1,1) coherent states The Perelomov coherent state of the su(1,1) algebra is defined as $`|k\text{ }\eta _P`$ $`=`$ $`S(\xi )|k\text{ }0`$ (39) $`=`$ $`\mathrm{exp}(\xi K_+\xi ^{}K_+)|k\text{ }0`$ (40) $`=`$ $`(1|\eta |^2)^k{\displaystyle \underset{n=0}{\overset{\mathrm{}}{}}}\sqrt{{\displaystyle \frac{\mathrm{\Gamma }(2k+n)}{\mathrm{\Gamma }(2k)n!}}}\eta ^n|k\text{ }n,`$ (41) where $`\xi =r\mathrm{exp}(i\theta )`$, $`\eta =\mathrm{exp}(i\theta )\mathrm{tanh}r`$, $`\mathrm{\Gamma }(x)`$ is the Gamma function and $`S(\xi )`$ is the su(1,1) displacement operator. We define a new SU(1,1) coherent state in association with the su(1,1) parity operator $`|k\text{ }\eta _{P\mathrm{\Pi }}`$ $`=`$ $`\stackrel{~}{S}(\xi )|k\text{ }0`$ (42) $`=`$ $`\mathrm{exp}(\xi \stackrel{~}{K}_+\xi ^{}\stackrel{~}{K}_{})|k\text{ }0,`$ (43) It is found that the Perelomov SU(1,1) coherent state is a nonlinear coherent state with the nonlinear function $`1/(𝒩+2k)`$. Therefore, the parity Perelomov SU(1,1) coherent state is a nonlinear coherent state with the nonlinear function$`(1)^𝒩/(𝒩+2k).`$ The normally-ordered su(1,1) displacement operator is $`\stackrel{~}{S}(\xi )`$ $`=`$ $`\mathrm{exp}(\eta \stackrel{~}{K}_+)(1|\eta |^2)^{\stackrel{~}{K}_z}\mathrm{exp}(\eta ^{}\stackrel{~}{K}_{}),`$ (44) $`\eta `$ $`=`$ $`{\displaystyle \frac{\xi }{|\xi |}}\mathrm{tanh}(|\xi |).`$ (45) Using the above equation, we obtain $`|k\text{ }\eta _{P\mathrm{\Pi }}`$ (46) $`=`$ $`{\displaystyle \frac{1}{\sqrt{2}}}\left(e^{i\frac{\pi }{4}}|k\text{ }i\eta _P+e^{i\frac{\pi }{4}}|k\text{ }i\eta _P\right).`$ (47) The parity Perelomov SU(1,1) coherent states are superpositions of two Perelomov SU(1,1) coherent states. The general entangled SU(1,1) coherent state is treated in Appendix A, but here we consider the two-particle case. From the two su(1,1) algebras for the two Hilbert space concerned with the entanglement, we define two new su(1,1) algebras as $`[\stackrel{~}{K}_+^n,\stackrel{~}{K}_{}^l]`$ $`=`$ $`2\delta _{nl}\stackrel{~}{K}_z^n,`$ (48) $`[\stackrel{~}{K}_z^n,\stackrel{~}{K}_\pm ^l]`$ $`=`$ $`\pm \delta _{nl}\stackrel{~}{K}_\pm ^l,\text{ }[\stackrel{~}{K}_{}^n,\stackrel{~}{K}_{}^l]=0.`$ (49) where $`\stackrel{~}{K}_+^n`$ $`=`$ $`K_+^n\mathrm{\Pi },\text{ }\stackrel{~}{K}_{}^n=\mathrm{\Pi }K_{}^n\text{}\stackrel{~}{K}_z^nK_z^n,`$ (50) $`\mathrm{\Pi }`$ $`=`$ $`(1)^{𝒩_1+𝒩_2},\text{ }n,l=1,2.`$ (51) The su(1,1) parity operator $`\mathrm{\Pi }`$ satisfies $`\mathrm{\Pi }^2=1`$ and $`\mathrm{\Pi }^{}=\mathrm{\Pi }`$. The Perelomov SU(1,1) coherent state of the two new su(1,1) algebras is obtained as $`|k\text{ }\stackrel{}{\eta }_{P\mathrm{\Pi }}`$ $`=`$ $`\stackrel{~}{S}(\stackrel{}{\xi })|k\text{ }\stackrel{}{0}`$ (52) $`=`$ $`{\displaystyle \frac{1}{\sqrt{2}}}\left(e^{i\frac{\pi }{4}}|k\text{ }i\stackrel{}{\eta }_P+e^{i\frac{\pi }{4}}|k\text{ }i\stackrel{}{\eta }_P\right)`$ (53) Here $`\stackrel{~}{S}(\stackrel{}{\xi })\stackrel{~}{S}_1(\xi _1)\stackrel{~}{S}_2(\xi _2).`$ The state $`|k`$ $`\stackrel{}{\eta }_{P\mathrm{\Pi }}`$ is the entangled Perelomov SU(1,1) coherent state. Again each Hilbert space is restricted to the same irrep $`k,`$ similar to the restriction for the su(2) case. #### 2 Entangled Barut-Girardello SU(1,1) coherent states There is another coherent state of the su(1,1) algebra known as the Barut-Girardello coherent state. It is defined as the eigenstate of the lowering operator $`K_{}`$ $$K_{}|k\text{ }\eta _{BG}=\eta |k\text{ }\eta _{BG},$$ (54) and it can be expressed as $$|k\text{ }\eta _{BG}=\sqrt{\frac{|\eta |^{2k1}}{I_{2k1}(2|\eta |)}}\underset{n=0}{\overset{\mathrm{}}{}}\frac{\eta ^n}{\sqrt{n!\mathrm{\Gamma }(n+2k)}}|k\text{ }n,$$ (55) where $`I_\nu (x)`$ is the modified Bessel function of the first kind. The Perelomov coherent state is defined with respect to the displacement operator formalism, whereas the Barut-Girardello coherent state is defined with respect to the ladder operator formalism. Thus, we define the parity Barut-Girardello SU(1,1) coherent state as $$\stackrel{~}{K}_{}|k\text{ }\eta _{BG\mathrm{\Pi }}=(1)^𝒩K_{}|k\text{ }\eta _{BG\mathrm{\Pi }}=\eta |k\text{ }\eta _{BG\mathrm{\Pi }}.$$ (56) The state $`|k`$ $`\eta _{BG\mathrm{\Pi }}`$ is a nonlinear coherent state with the nonlinear function $`(1)^𝒩`$. From the general expression of a SU(1,1) nonlinear coherent state, we obtain the expression of the state $`|k`$ $`\eta _{BG\mathrm{\Pi }}`$ as $$|k\text{ }\eta _{BG\mathrm{\Pi }}=\frac{1}{\sqrt{2}}\left(e^{i\frac{\pi }{4}}|k\text{ }i\eta _{BG}+e^{i\frac{\pi }{4}}|k\text{ }i\eta _{BG}\right).$$ (57) Eigenstates of operators $`\stackrel{~}{K}_{}^n(n=1,2)`$ are constructed as $`|k\text{ }\stackrel{}{\eta }_{BG\mathrm{\Pi }}`$ $``$ $`\mathrm{exp}\left({\displaystyle \frac{\eta _1}{𝒩_1+2k1}}\stackrel{~}{K}_+^1\right)`$ (59) $`\mathrm{exp}\left({\displaystyle \frac{\eta _2}{𝒩_2+2k1}}\stackrel{~}{K}_+^2\right)|k\text{ }\stackrel{}{0}`$ $`=`$ $`{\displaystyle \frac{1}{\sqrt{2}}}\left(e^{i\frac{\pi }{4}}|k\text{ }i\stackrel{}{\eta }_{BG}+e^{i\frac{\pi }{4}}|k\text{ }i\stackrel{}{\eta }_{BG}\right),`$ (60) which is the entangled Barut-Girardello SU(1,1) coherent states. Here we have used the exponential form of the unnormalized Barut-Girardello coherent state $$|k\text{ }\eta _{BG}\mathrm{exp}\left(\frac{\eta }{(𝒩+2k1)}K_+\right)|k\text{ }0.$$ (61) The generalization of (2.19), (2.30) and (2.35) to multiparticle entangled SU(2) and SU(1,1) coherent states is treated in Appendix A. ## III Nonlinear su(2) and su(1,1) coherent states ### A SU(2) case The su(2) parity operator $`\mathrm{\Pi }=(1)^{}`$ is a special case of the unitary operator $`U()=\mathrm{exp}[i\vartheta ()])`$. Here $`\vartheta ()`$ is a nonsingular function of $`.`$ It is easy to check that the operators $$\overline{J}_+=J_+U(),\text{ }\overline{J}_{}=U^{}()J_{}\text{}\overline{J}_zJ_z$$ (62) satisfy the su(2) commutation relations (2.7). Then we can define the SU(2) coherent state corresponding to this su(2) algebra as $`|\vartheta ;\text{ }j\text{ }\gamma \text{ }`$ $``$ $`\overline{R}(\gamma )|jj`$ (63) $`=`$ $`\mathrm{exp}\left[{\displaystyle \frac{1}{2}}\theta \left(\overline{J}_+e^{i\varphi }\overline{J}_{}e^{i\varphi }\right)\right]|jj.`$ (64) By use of (2.14), the state $`|\vartheta ;`$ $`j`$ $`\gamma `$ is obtained as $`|\vartheta ;\text{ }j\text{ }\gamma \text{ }=\left(1+|\gamma |^2\right)^j{\displaystyle \underset{m=0}{\overset{2j}{}}}\left({\displaystyle \genfrac{}{}{0pt}{}{2j}{m}}\right)^{1/2}\gamma ^m`$ (65) $`\times \mathrm{exp}\left[i{\displaystyle \underset{n=1}{\overset{m}{}}}\vartheta (2jn)\right]|jjm,`$ (66) In the derivation of the above equation, we have used the relation $$[f()J_{}]^m=(J_{})^m\underset{n=1}{\overset{m}{}}f(n)$$ (67) If we choose $`\vartheta ()=\pi ,`$ equation (3.3) reduces to (2.15) as we expected. We refer to the state $`|\vartheta ;`$ $`j`$ $`\gamma `$ $``$ as a nonlinear SU(2) coherent state if $`\vartheta ()`$ is a nonlinear function of $``$. ### B SU(1,1) case By analogy to the su(2) case (3.1), we define $$\overline{K}_+=K_+V(𝒩),\text{ }\overline{K}_{}=V^{}(𝒩)K_{}\text{}\overline{K}_zK_z.$$ (68) where $`V(𝒩)=\mathrm{exp}[i\phi (𝒩)]`$. These operators satisfy the su(1,1) commutation relations (2.20). Then we can define the Perelomov SU(1,1) coherent state corresponding to this su(1,1) algebra as $`|\phi ;\text{ }k\text{ }\eta _P`$ $`=`$ $`\overline{S}(\xi )|k\text{ }0`$ (69) $`=`$ $`\mathrm{exp}(\xi \overline{K}_+\xi ^{}\overline{K}_{})|k\text{ }0.`$ (70) As is mentioned in the last section, the Perelomov SU(1,1) coherent state $`|k`$ $`\eta _P`$ is a nonlinear coherent state with the nonlinear function $`1/(𝒩+2k).`$ Therefore, the state $`|\phi ;`$ $`k`$ $`\eta _P`$ is a nonlinear coherent state with the nonlinear function $`\mathrm{exp}[i\phi (𝒩)]/(𝒩+2k).`$ From the general expression for an su(1,1) nonlinear coherent, we obtain $`|\phi ;\text{ }k\text{ }\eta _P`$ $`=`$ $`(1|\eta |^2)^k{\displaystyle \underset{n=0}{\overset{\mathrm{}}{}}}\sqrt{{\displaystyle \frac{\mathrm{\Gamma }(2k+n)}{\mathrm{\Gamma }(2k)n!}}}\eta ^n`$ (72) $`\times \mathrm{exp}\left[i{\displaystyle \underset{m=0}{\overset{n1}{}}}\phi (m)\right]|k\text{ }n.`$ In the derivation of the above equation, we have used the relation $$[K_+f(𝒩)]^m=(K_+)^m\underset{n=1}{\overset{m}{}}f(𝒩+n1).$$ (73) The nonlinear Barut-Girardello coherent state is defined as $`\overline{K}_{}|\phi ;\text{ }k\text{ }\eta _{BG}`$ $`=`$ $`\mathrm{exp}[i\phi (𝒩)]K_{}|\phi ;\text{ }k\text{ }\eta _{BG}`$ (74) $`=`$ $`\eta |\phi ;\text{ }k\text{ }\eta _{BG}.`$ (75) The state $`|\phi ;`$ $`k`$ $`\eta _{BG}`$ is also a nonlinear coherent state with the nonlinear function $`\mathrm{exp}[i\phi (𝒩)].`$ The expansion of this state yields $`|\phi ;\text{ }k\text{ }\eta _{BG}`$ $`=`$ $`\sqrt{{\displaystyle \frac{|\eta |^{2k1}}{I_{2k1}(2|\eta |)}}}{\displaystyle \underset{n=0}{\overset{\mathrm{}}{}}}{\displaystyle \frac{1}{\sqrt{\mathrm{\Gamma }(2k+n)n!}}}\eta ^n`$ (77) $`\times \mathrm{exp}[i{\displaystyle \underset{m=0}{\overset{n1}{}}}\phi (m)]|k\text{ }n.`$ If we choose $`\phi (𝒩)=\pi 𝒩,`$ equations (3.7) and (3.10) reduce to (2.27) and (2.34), respectively, as we expected. ## IV Representation of entangled su(2) and su(1,1) coherent states in Fock space ### A Entangled binomial states It is well known that the operators $$J_+=a^{}\sqrt{MN}\text{}J_{}=\sqrt{MN}a,\text{ }J_z=NM/2$$ (78) generate the su(2) algebra via the Holstein-Primakoff representation in the spin $`M/2`$ representation. Here $`N=a^{}a`$ is the number operator. The vacuum $`|0`$ is the lowest weight state $$J_{}|0=0,\text{ }J_z|0=M/2|0.$$ (79) Then we define the SU(2) coherent state in the displacement operator form $`|M\text{ }\eta `$ $`=`$ $`\mathrm{exp}(\zeta J_+\zeta ^{}J_{})|0`$ (80) $`=`$ $`\mathrm{exp}(\zeta a^{}\sqrt{MN}\zeta ^{}\sqrt{MN}a)|0`$ (81) where we choose $`\zeta =\mathrm{exp}(i\theta )\mathrm{arctan}(|\eta |/\sqrt{1|\eta |^2}),`$ $`\eta =|\eta |\mathrm{exp}(i\theta ).`$ The normally ordered form of the displacement operator is $`\mathrm{exp}(\zeta J_+\zeta ^{}J_{})`$ (82) $`=\mathrm{exp}(\stackrel{~}{\eta }J_+)(1+|\stackrel{~}{\eta }|^2)^{J_z}\mathrm{exp}(\stackrel{~}{\eta }^{}J_{})`$ (83) where $`\stackrel{~}{\eta }=\zeta /|\zeta |\mathrm{tan}|\zeta |=\eta /\sqrt{1|\eta |^2}.`$ By use of (4.4), we obtain $$|M\text{ }\eta =(1|\eta |^2)^{M/2}\underset{n=0}{\overset{M}{}}\left(\genfrac{}{}{0pt}{}{M}{n}\right)^{1/2}\left(\frac{\eta }{\sqrt{1|\eta |^2}}\right)^n|n.$$ (84) This is the binomial state. Therefore, the binomial state is a special type of SU(2) coherent state via the Holstein-Primakoff representation. From (2.8) and (4.1) we see that the parity operator is $`(1)^N.`$ By analogy to the discussion in Section II, we can obtain the entangled binomial states of the type (2.19). Now we show how to generate entangled binomial states in a particular Hamiltonian system. The entangled coherent states can be created using an ideal Kerr nonlinearity with three nonlinear media elements. We show that the entangled binomial state can also be generated in this system. By an appropriate arrangement of the three nonlinear elements, the effective Kerr transformation with two input fields, 1 and 2, is given by $$S_{12}(\chi )=\mathrm{exp}(i\chi a_1^{}a_1a_2^{}a_2).$$ (85) We assume that the initial state is a product of binomial states $`|M`$ $`\stackrel{}{\eta }=|M`$ $`\eta _1_1|M`$ $`\eta _2_2.`$ For $`\chi =\pi ,`$ the resulting output state is $`{\displaystyle \frac{1}{2}}(|M\text{ }\eta _1_1+|M\text{ }\eta _1_1)|M\text{ }\eta _2_2`$ (86) $`+(|M\text{ }\eta _1_1|M\text{ }\eta _1_1)|M\text{ }\eta _2_2`$ (87) This state contains both even and odd binomial states. ### B Entangled negative binomial states The generators of the su(1,1) algebra via the Holstein-Primakoff realization of the discrete irreducible representation with Bargmann index $`M/2`$ are $$K_+=a^{}\sqrt{M+N}\text{}K_{}=\sqrt{M+N}a,\text{ }K_z=N+M/2,$$ (88) and the vacuum $`|0`$ is the lowest-weight state: $$K_{}|0=0,\text{ }K_z|0=M/2|0.$$ (89) The corresponding SU(1,1) coherent state is $$|M\text{ }\eta ^{}=\mathrm{exp}(\xi K_+\xi ^{}K_{})|0.$$ (90) Here $`\xi =\mathrm{exp}(i\theta )\mathrm{arctan}`$h$`|\eta |.`$ Using (2.26), we obtain $$|M\text{ }\eta ^{}=(1|\eta |^2)^{M/2}\underset{n=0}{\overset{\mathrm{}}{}}\left(\genfrac{}{}{0pt}{}{M+n1}{n}\right)^{1/2}\eta ^n|n.$$ (91) This is the negative binomial state. Therefore, the negative binomial state is a special type of SU(1,1) Perelomov coherent state. The parity operator is $`(1)^N`$ and the entangled negative binomial states of the type (2.30) can be obtained. We do not write these entangled states explicitly here. Under the transformation $`S_{12}(\pi )`$, the input negative binomial state $`|M`$ $`\stackrel{}{\eta }^{}`$ will be transformed into the entangled negative binomial states $`{\displaystyle \frac{1}{2}}(|M\text{ }\eta _1_1^{}+|M\text{ }\eta _1_1^{})|M\text{ }\eta _2_2^{}`$ (92) $`+(|M\text{ }\eta _1_1^{}|M\text{ }\eta _1_1^{})|M\text{ }\eta _2_2^{}.`$ (93) This state contains both even and odd negative binomial states ### C Contraction of su(2) and su(1,1) algebra Let us note the fact that the binomial distribution tends to the Poisson distribution in a certain limit. Let $`M\mathrm{},|\eta |0`$ in such a way that the product $`|\eta |^2M=|\alpha |^2`$ is fixed. In this limit, the binomial distribution of the binomial state tends to the Poisson distribution $`\mathrm{exp}(|\alpha |^2)|\alpha |^{2n}/n!,`$ and the binomial state tends to the ordinary coherent state $$|M\text{ }\eta \mathrm{exp}(|\alpha |^2/2)\underset{n=0}{\overset{\mathrm{}}{}}\frac{\alpha ^n}{\sqrt{n!}}|n.$$ (94) Here we have used the relation $$(1|\eta |^2)^M\mathrm{exp}(|\alpha |^2).$$ (95) This limit can also be visualized as a contraction of the su(2) algebra into the Heisenberg-Weyl algebra $$|\eta |J_+|\alpha |a^{},|\eta |J_{}|\alpha |a.$$ (96) Thus, (4.5) tends to the coherent state $$|M\text{ }\eta \mathrm{exp}[\alpha a^{}\alpha ^{}a]|0.$$ (97) In this limit the entangled binomial state (4.7) reduces to the entangled coherent state $`{\displaystyle \frac{1}{2}}[(|\alpha _1_1+|\alpha _1_1)|\alpha _2_2`$ (98) $`+(|\alpha _1_1|\alpha _1_1)|\alpha _2_2`$ (99) which can be employed as qubits in quantum computation. In the same limit described above, the negative binomial states reduce to coherent states, the su(1,1) algebra contracts into the Heisenberg-Weyl algebra, and the entangled negative binomial states (4.12) reduce to the entangled harmonic oscillator coherent states (4.17). ### D Entangled squeezed states The amplitude-squared su(1,1) algebra is realized by $$K_+=\frac{1}{2}a^{+2},K_{}=\frac{1}{2}a^2,K_z=\frac{1}{2}\left(N+\frac{1}{2}\right).$$ (100) The representation on the usual Fock space is completely reducible and decomposes into a direct sum of the even Fock space ($`S_0`$) and odd Fock space ($`S_1`$), $$S_i=\text{span}\{|n_i|2n+i|n=0,1,2,\mathrm{}\},\text{ }i=0,1.$$ (101) Representations on $`S_i`$ can be written as $`K_+|n_i`$ $`=`$ $`\sqrt{(n+1)(n+i+1/2)}|n+1_i,`$ (102) $`K_{}|n_i`$ $`=`$ $`\sqrt{(n)(n+i1/2)}|n1_i,`$ (103) $`K_0|n_i`$ $`=`$ $`(n+i/2+1/4)|n_i.`$ (104) The Bargmann index $`k=1/4`$ ($`3/4`$) for even (odd) Fock space. We see that the Perelomov SU(1,1) coherent states in even/odd Fock space are squeezed vacuum states and squeezed first Fock states $`|\eta _V`$ $`=`$ $`\mathrm{exp}\left({\displaystyle \frac{\xi }{2}}a^{+2}{\displaystyle \frac{\xi ^{}}{2}}a^2\right)|0,`$ (105) $`|\eta _F`$ $`=`$ $`\mathrm{exp}\left({\displaystyle \frac{\xi }{2}}a^{+2}{\displaystyle \frac{\xi ^{}}{2}}a^2\right)|1,`$ (106) respectively. From (2.21) and (4.18), the corresponding parity operator is $`(1)^{N/2}`$ in even Fock space and $`(1)^{(N1)/2}`$ in odd Fock space. We consider two modes $`a_1`$ and $`a_2.`$ Then from (2.30) and through the amplitude-squared su(1,1) realization, we obtain the entangled squeezed vacuum states and entangled squeezed first Fock states as $`|\stackrel{}{\eta }_V={\displaystyle \frac{1}{\sqrt{2}}}\left(e^{i\frac{\pi }{4}}|i\stackrel{}{\eta }_V+e^{i\frac{\pi }{4}}|i\stackrel{}{\eta }_V\right),`$ (107) $`|\stackrel{}{\eta }_F={\displaystyle \frac{1}{\sqrt{2}}}\left(e^{i\frac{\pi }{4}}|i\stackrel{}{\eta }_F+e^{i\frac{\pi }{4}}|i\stackrel{}{\eta }_F\right).`$ (108) The state (4.22) is a special case of the entangled squeezed coherent state for zero amplitude and reduces to the superposition of two squeezed vacuum states. In Fock space, we have obtained entangled binomial states, entangled negative binomial states and entangled squeezed states. They are special cases of entangled SU(2) coherent states or entangled SU(1,1) coherent states. ## V Generation of the entangled coherent states A superposition of two distinct su(2) coherent states can be generated by the Hamiltonian system for the nonlinear rotator $$H_j=\omega J_z+\frac{\lambda }{2j}J_z^2\text{ (}\mathrm{}=1\text{)}$$ (109) where $`\omega `$ is the linear precession frequency and $`\lambda `$ is a positive constant. Let the initial state be the SU(2) coherent state $`|j,\gamma .`$ At time $`t_j=\pi j/\lambda `$, the coherent state has evolved into the superposition state $`\mathrm{exp}(iH_jt_j)|j\text{ }\gamma `$ (110) $`=`$ $`{\displaystyle \frac{1}{\sqrt{2}}}\left(e^{i\frac{\pi }{4}}|j\text{ }\overline{\gamma }+e^{i\frac{\pi }{4}}(1)^j|j\text{ }\overline{\gamma }\right).`$ (111) Here $`\overline{\gamma }=\mathrm{exp}(i\omega \pi j/\lambda )\gamma .`$ A superposition of two coherent states, separated by a phase $`\pi ,`$ has been generated from the initial SU(2) coherent state. The phase difference of the coefficients in (5.2) depends on whether $`j`$ is odd or even. Let $`\overline{\gamma }i(1)^{2j}\gamma `$ and $`j`$ be even; the above state is just the parity SU(2) coherent state $`|j`$ $`\gamma _\mathrm{\Pi }.`$ Gerry and Bu$`\stackrel{ˇ}{z}`$ek have studied the dynamics of the nonlinear oscillator with the Hamiltonian $`H`$ written as $$H=\omega a^{}a+\frac{\lambda }{2}a^2a^2=\omega K_z+\lambda K_+K_{}$$ (112) up to constant terms. Here $`K_z`$ and $`K_\pm `$ are generators of the amplitude-squared su(1,1) algebra (4.18). At time $`t=\pi /2\lambda `$, the initial coherent state has evolved into the superposition state $`{\displaystyle \frac{1}{\sqrt{2}}}\left[e^{i\frac{\pi }{4}}|k\text{ }\overline{\eta }_P+e^{i\frac{\pi }{4}}|k\text{ }\overline{\eta }_P\right],`$ (113) $`{\displaystyle \frac{1}{\sqrt{2}}}\left[e^{i\frac{\pi }{4}}|k\text{ }\overline{\eta }_{BG}+e^{i\frac{\pi }{4}}|k\text{ }\overline{\eta }_{BG}\right],`$ (114) where $`\overline{\eta }=\mathrm{exp}\{i\pi [\omega +(2k1)\lambda ]/(2\lambda )\}.`$ Here we choose the initial state to be the SU(1,1) coherent state $`|k`$ $`\eta _P`$ and $`|k`$ $`\eta _{BG}`$, respectively. Let $`\overline{\eta }i\eta ;`$ the above superposition states are then just the parity coherent states $`|k\eta _{P\mathrm{\Pi }}`$ and $`|k`$ $`\eta _{BG\mathrm{\Pi }}`$. Now we consider how to generate the entangled SU(2) and SU(1,1) coherent states. By analogy with the generation of entangled coherent states of the harmonic oscillator, the two Hamiltonians are given by $`H_2`$ $`=`$ $`\chi _1_1^2+\chi _2_2^2+\chi _3_1_2,`$ (115) $`H_{11}`$ $`=`$ $`\lambda _1𝒩_1^2+\lambda _2𝒩_2^2+\lambda _3𝒩_1𝒩_2.`$ (116) We assume that the initial state of the Hamiltonian system $`H_2`$ is $`|j`$ $`i\stackrel{}{\gamma }`$ and the time $`t=\pi /(2\chi _1),\chi _2=\chi _1,\chi _3=2\chi _1`$. Then the initial state evolves into the entangled SU(2) coherent state $`|j`$ $`\stackrel{}{\gamma }_\mathrm{\Pi }`$ $`(j`$ is an integer$`).`$ Similarly, we assume that the initial state of the Hamiltonian system $`H_{11}`$ is $`|k`$ $`i\stackrel{}{\eta }_P`$ and the time $`t=\pi /(2\lambda _1),`$ $`\lambda _2=\lambda _1,`$ $`\lambda _3=2\lambda _1`$. The initial state evolves into the entangled SU(1,1) coherent state $`|k`$ $`\stackrel{}{\eta }_{P\mathrm{\Pi }\text{.}}`$ If we choose the initial state as an SU(1,1) coherent state $`|k`$ $`i\stackrel{}{\eta }_{BG},`$ the resulting state will be the entangled SU(1,1) coherent states $`|k`$ $`\stackrel{}{\eta }_{BG\mathrm{\Pi }\text{.}}`$ It is interesting if we choose the parameters $`\chi _1=\chi _2=0,t=\pi /\chi _3.`$ The initial state $`|j\stackrel{}{\gamma }`$ will evolve into the following entangled SU(2) coherent state $`{\displaystyle \frac{1}{2}}\{[|j\text{ }\gamma _1_1+(1)^{2j}|j\text{ }\gamma _1_1]|j\text{ }\gamma _2_2+`$ (117) $`[(1)^{2j}|j\text{ }\gamma _1_1|j\text{ }\gamma _1_1]|j\text{ }\gamma _2_2\}`$ (118) $`={\displaystyle \frac{1}{2}}\{|j\text{ }\gamma _1_1[|j\text{ }\gamma _2_2+(1)^{2j}|j\text{ -}\gamma _2_2]+`$ (119) $`|j\text{ }\gamma _1_1[(1)^{2j}|j\text{ }\gamma _2_2|j\text{ }\gamma _2_2]\}.`$ (120) The entangled coherent state (5.8) includes the SU(2) ‘cat’ states. We assume that the initial state of the Hamiltonian system $`H_{11}`$ is $`|k`$ $`\stackrel{}{\eta }.`$ Here the subscripts $`P`$ and $`BG`$ have been omitted since the discussions are the same for the Perelomov coherent states and the Barut-Girardello coherent states. At time $`t=\pi /\lambda _3`$ and $`\lambda _1=\lambda _2=0,`$ Resultant state is $`{\displaystyle \frac{1}{2}}[(|k\text{ }\eta _1_1+|k\text{ }\eta _1_1)|k\text{ }\eta _2_2+`$ (121) $`(|k\text{ }\eta _1_1|k\text{ }\eta _1_1)|k\text{ }\eta _2_2],`$ (122) which includes the SU(1,1) ‘cat’ states. ## VI Degree of entanglement ### A Bell inequality A standard example of entanglement of two-particle nonorthogonal states is given by $$|\mathrm{\Psi }=\mu |\alpha _1|\beta _2+\nu |\gamma _1|\delta _2,$$ (123) where $`|\alpha _1`$ and $`|\gamma _1`$ are nonorthogonal states of system 1 and similarly for $`|\beta _2`$ and $`|\delta _2`$ of system 2. We also assume that the states $`|\alpha _1`$ and $`|\gamma _1`$ are linearly independent as are $`|\beta _2`$ and $`|\delta _2`$. The entangled SU(2) and SU(1,1) coherent states obtained in the previous section are special cases of these entangled nonorthogonal states $`|\mathrm{\Psi }`$. The state $`|\mathrm{\Psi }`$ is a pure state whose density matrix is $`\rho _{12}=|\mathrm{\Psi }\mathrm{\Psi }|.`$ Then the reduced density matrices $`\rho _1=`$Tr$`{}_{2}{}^{}(\rho _{12})`$ and $`\rho _2=`$Tr$`{}_{1}{}^{}(\rho _{12})`$ for each subsystem can be obtained, and the two eigenvalues of $`\rho _1`$ are given by $`\lambda _\pm `$ $`=`$ $`{\displaystyle \frac{1}{2}}\pm {\displaystyle \frac{1}{2}}\sqrt{14|\mu \nu 𝒜_1𝒜_2|^2},`$ (124) $`𝒜_1`$ $`=`$ $`\sqrt{1|_1\alpha |\gamma _1|^2},\text{ }𝒜_2=\sqrt{1|_2\delta |\beta _2|^2}.`$ (125) The two eigenvalues of $`\rho _{2\text{ }}`$are identical to those of $`\rho _{1\text{.}}`$ The corresponding eigenvectors of $`\rho _1`$ are $`|\pm _1,`$ and the corresponding orthonormal eigenvectors of $`\rho _2`$ are denoted by $`|\pm _2.`$ The general theory of the Schmidt decomposition implies that the state $`|\mathrm{\Psi }`$ can be expressed in the Schmidt form $$|\mathrm{\Psi }=c_{}|_1|_2+c_+|+_1|+_2,$$ (126) where $`|c_\pm |^2=\lambda _\pm ,`$ $`|c_+|^2+|c_{}|^2=1.`$ The two-system entangled state(6.3) violates a Bell inequality. More specifically, we choose Hermitian operator $`\widehat{\mathrm{\Theta }}_i`$ for each subsystem such that the eigenvalues are $`\pm 1.`$ The general form for such an operator is $`\widehat{\mathrm{\Theta }}_i(\lambda _i,\phi _i)`$ $`=`$ $`\mathrm{cos}\lambda _i(|+_i+||_i|)`$ (128) $`+\mathrm{sin}\lambda _i(e^{i\phi _i}|+_i|+e^{i\phi _i}|_i+|).`$ The Bell operator is defined as $$\widehat{B}=\widehat{\mathrm{\Theta }}_1\widehat{\mathrm{\Theta }}_2+\widehat{\mathrm{\Theta }}_1\widehat{\mathrm{\Theta }}_2^{}+\widehat{\mathrm{\Theta }}_1^{}\widehat{\mathrm{\Theta }}_2\widehat{\mathrm{\Theta }}_1^{}\widehat{\mathrm{\Theta }}_2^{},$$ (129) for $`\widehat{\mathrm{\Theta }}_i\widehat{\mathrm{\Theta }}_i(\lambda _i,\phi _i),`$ $`\widehat{\mathrm{\Theta }}_i^{}\widehat{\mathrm{\Theta }}_i(\lambda _i^{},\phi _i^{}).`$ For the choices $`\lambda _1`$ $`=`$ $`0,\text{ }\lambda _1^{}=\pi /2,`$ (130) $`\lambda _2`$ $`=`$ $`\text{ }\lambda _2^{}=\mathrm{cos}^1[1+|2c_+c_{}|^2]^{1/2},`$ (131) $`\phi _1+\phi _2`$ $`=`$ $`\phi _1^{}+\phi _2^{}=\phi _+\phi _{},`$ (132) where $`\phi _\pm `$ are the phases of $`c_\pm `$ in (6.3), the expectation value of the Bell operator for the state $`|\mathrm{\Psi }`$ $`(6.1)`$ is $$B=\mathrm{\Psi }|\widehat{B}|\mathrm{\Psi }=2\sqrt{1+4\lambda _+\lambda _{}}>2.$$ (133) The degree of violation depends on the values of $`\lambda _\pm ,`$ but a violation always occurs provided that the state is entangled. Several entangled coherent states are obtained in the previous sections. Here we only consider two examples. The first example is the entangled SU(2) coherent state $`|j`$ $`\stackrel{}{\gamma }_\mathrm{\Pi }`$ (2.19). For simplicity $`\gamma _0=\gamma _1=\gamma _2`$ is introduced. The eigenvalues of the reduced density matrices $`\rho _1`$ and $`\rho _2`$ are $$\lambda _\pm =\frac{1}{2}\pm \frac{1}{2}\sqrt{1\left[1\left(\frac{1|\gamma _0|^2}{1+|\gamma _0|^2}\right)^{4j}\right]^2}.$$ (134) Then the expectation value of the Bell operator for the entangled coherent state $`|j`$ $`\stackrel{}{\gamma }_\mathrm{\Pi }`$ is given by $$B=2\sqrt{1+\left[1\left(\frac{1|\gamma _0|^2}{1+|\gamma _0|^2}\right)^{4j}\right]^2}.$$ (135) The second example is the entangled Perelomov SU(1,1) coherent state $`|k\stackrel{}{\eta }_{P\mathrm{\Pi }}`$. The corresponding eigenvalues and expectation value of the Bell operator are obtained as $`\lambda _\pm `$ $`=`$ $`{\displaystyle \frac{1}{2}}\pm {\displaystyle \frac{1}{2}}\sqrt{1\left[1\left({\displaystyle \frac{1|\eta _0|^2}{1+|\eta _0|^2}}\right)^{4k}\right]^2},`$ (136) $`B`$ $`=`$ $`2\sqrt{1+\left[1\left({\displaystyle \frac{1|\eta _0|^2}{1+|\eta _0|^2}}\right)^{4k}\right]^2}.`$ (137) Here we have chosen $`\eta _0=\eta _1=\eta _2.`$ ### B Entropy The entropy $`S`$ of a quantum state described by density operator $`\rho `$ is defined by $$S=k_B\text{Tr(}\rho \mathrm{ln}\rho ),$$ (138) where $`k_B`$ is the Boltzmann constant. The entropy defined above is zero for a pure state and positive for a mixed state. If we consider two systems, the entropy the system 1(2) is determined via the reduced density operator $$S_{1(2)}=k_B\text{Tr}_{1(2)}\text{(}\rho _{1(2)}\mathrm{ln}\rho _{1(2)}).$$ (139) In 1970 Araki and Lieb proved the following inequality: $$|S_1S_2|SS_1+S_2.$$ (140) One consequence of this inequality is that, if the total system is in a pure state, $`S_1=S_2.`$ From an information theory point of view, the entropy can be regarded as the amount of uncertainty contained within the density operator. We can use the index of correlation $`I_c`$ as the amount of information lost in the tracing procedure $$I_c=S_1+S_2S.$$ (141) For the pure state $`|\mathrm{\Psi },`$ the index of correlation is obtained as $$I_c=2S_1=2k_B(\lambda _+\mathrm{ln}\lambda _++\lambda _{}\mathrm{ln}\lambda _{}).$$ (142) If $`I_c=0,`$ the two subsystems are in a pure state and disentangled. The combination of (6.8-6.11) and (6.16) gives the index of correlation of entangled SU(2) and SU(1,1) coherent states. ### C Discussion We now discuss two limit cases for certain parameters. When $`|`$$`\gamma _0|(|\eta _0|)0`$ or $`\mathrm{},`$ $`B=2,`$ i.e., the entangled coherent states become product states and disentangled. When $`|`$$`\gamma _0|(|\eta _0|)=1,`$ $`B=2\sqrt{2},`$ and the entangled coherent states are maximally entangled states. For this case, the two states $`|j`$ $`\gamma _0`$ and $`|j`$ $`\gamma _0`$ are orthogonal and the states $`|k`$ $`\eta _0`$ and $`|k`$ $`\eta _0`$ become orthogonal. When $`|`$$`\gamma _0|(|\eta _0|)0`$ or $`\mathrm{},`$ $`\lambda _+=1`$ and $`\lambda _{}=0,`$ then the index of correlation is zero as we expect. The entangled states become product states. When $`|`$$`\gamma _0|(|\eta _0|)=1,`$ $`\lambda _\pm =1/2,`$ then the index of correlation is $`2k_B\mathrm{ln}2,`$ which results from the orthogonality of the states $`|j`$ $`\pm \gamma _0(|\gamma _0|=1)`$ or $`|k`$ $`\pm \eta _0(|\eta _0|=1).`$ An index of correlation is maximum for $`I_c=2k_B\mathrm{ln}2`$ . The maximum entanglement of SU(2) coherent states $`|j`$ $`\stackrel{}{\gamma }_\mathrm{\Pi }`$ for $`\stackrel{}{\gamma }=(\gamma _0,\gamma _0)`$ and $`|\gamma _0|=1`$ can be understood as follows. From the expression for $`\gamma =\mathrm{exp}(i\varphi )\mathrm{tan}(\theta /2),`$ we can see that $`|\gamma |=1`$ corresponds to the set of all points on the equator ($`\theta =\pi /2`$) of the Poincar$`\stackrel{´}{e}`$ sphere, which is used to represent the pure states of an SU(2) system. The states corresponding to $`|j`$ $`\gamma _0=1`$ and $`|j\gamma _0=1`$ are antipodal; that is, one state is represented by a point at the intersection of the equator and the longitudinal line $`\varphi =0,`$ and the other state is represented by the point at the intersection of the equator and the longitudinal line $`\varphi =\pi .`$ We can thus employ the notation of Eq. (1.1), but with an appropriate SU(2) rotation (and allow for arbitrary $`j`$) by replacing $`|j`$ $`\gamma _0=1`$ by $`|1`$ and $`|j`$ $`\gamma _0=1`$ by $`|0.`$ For $`j=1/2,`$ we recover the qubit case (1.1). The entanglement state for $`j=1/2`$ is just $`2^{1/2}\left[\mathrm{exp}(i\pi /4)|11+\mathrm{exp}(i\pi /4)|00\right].`$ This state is a form of Bell state (1.3) and the reason for $`B`$ and $`I_c`$ being maximal is evident. A similar analysis can be employed for the entangled SU(1,1) coherent states represented by points on the Lobachevsky plane. ## VII Conclusions We have introduced entangled SU(2) and SU(1,1) coherent states. The general form for these entangled coherent states, which also incorporates entangled harmonic oscillator coherent states in the formalism, is given in the Appendix, but the main concern here is with two-particle coherent states. The two-particle coherent states present a diverse range of interesting phenomena, as we have shown. Aside from the mathematical elegance of entangled SU(2) and SU(1,1) coherent states, we have also applied these states to current topics of research, namely quantum information (specifically qubits) and nonclassical states of light (specifically squeezed vacuum states). We have explored several aspects of coherent states. Aside from defining these states, we have employed the entangled binomial states to establish a rigorous contraction from entangled SU(2) coherent states to entangled harmonic oscillator coherent states. Two measures of entanglement, the Bell operator approach and the index of correlation, have been used to quantify the degree of entanglement. As the entanglement is generally between non-orthogonal states, the degree of entanglement ranges from no entanglement (product state) to being maximally entangled for various parametric choices. The generation of entangled coherent states have been treated here by a Hamiltonian evolution which is a generalization of $`J_z^2`$ and $`K_z^2`$ nonlinear evolution for multiparticle systems. However, such evolutions are extremely sensitive to environmental-induced decoherence. Other methods for generating entangled coherent states could be considered, but the nonlinear evolution considered here illustrated one possible approach to producing these entangled states. These results can be generalized in various ways which would be of interest. One generalization is to entangled generalized coherent states, with generalized coherent states of the type treated by Perelomov. Another intriguing generalization is to entangled SU(2) and SU(1,1) coherent states for the respective Hilbert spaces, not restricted to the same irreducible representations. For example one could consider entanglement between spin-1/2 states (qubits) and spin-1 states (qutrits). Or one could consider entanglement of SU(1,1) coherent states with $`k=1/4`$ (single mode squeezed vacuum states) and SU(1,1) coherent states with $`k=3/4`$ (two-mode squeezed vacuum states). These ideas are topics for future research. ## A Most General form of Entangled SU(2) AND SU(1,1) coherent states The general form of entangled SU(2) and SU(1,1) coherent states can be written as $$𝑑\mu (\stackrel{}{\xi })f(\stackrel{}{\xi })|l\text{ }\stackrel{}{\xi },$$ (A1) where $`l=j`$ and $`\stackrel{}{\xi }=\stackrel{}{\gamma }`$ for SU(2) coherent states, and $`l=k`$ and $`\stackrel{}{\xi }=\stackrel{}{\eta }`$ for either Perelomov or Barut-Girardello SU(1,1) coherent states. We can also obtain the entangled coherent state (2.5) via the same expression by replacing $`\stackrel{}{\xi }`$ by $`\stackrel{}{\alpha }`$ and ignoring the irrep index $`l.`$ The measure $`d\mu (\stackrel{}{\xi })`$ is, for each case, $`d\mu (\stackrel{}{\gamma })`$ $`=`$ $`{\displaystyle \underset{n}{}}{\displaystyle \frac{2j+1}{\pi }}{\displaystyle \frac{d^2\stackrel{}{\gamma }_n}{(1+|\gamma _n|^2)^2}},`$ (A2) $`d\mu _P(\stackrel{}{\eta })`$ $`=`$ $`{\displaystyle \underset{n}{}}{\displaystyle \frac{2k1}{\pi }}{\displaystyle \frac{d^2\stackrel{}{\eta }_n}{(1|\eta _n|^2)^2}},`$ (A3) $`d\mu _{BG}(\stackrel{}{\eta })`$ $`=`$ $`{\displaystyle \underset{n}{}}{\displaystyle \frac{2}{\pi }}I_{2k1}^2(2|\eta _n|)\text{ }d^2\stackrel{}{\eta }_n,`$ (A4) $`d\mu (\stackrel{}{\alpha })`$ $`=`$ $`{\displaystyle \underset{n}{}}{\displaystyle \frac{d^2\stackrel{}{\alpha }}{\pi }}`$ (A5) for SU(2) coherent states, Perelomov SU(1,1) coherent states and Barut-Girardello coherent states, and harmonic oscillator coherent states, respectively. If we choose $`l=j`$ and $$f(\stackrel{}{\xi })=\frac{1}{\sqrt{2}}\left[\mathrm{exp}\left(i\frac{\pi }{4}\right)\delta (\stackrel{}{\xi }+i\stackrel{}{\gamma })+\mathrm{exp}\left(i\frac{\pi }{4}\right)\delta (\stackrel{}{\xi }i\stackrel{}{\gamma })\right],$$ (A6) the general state (A1) will reduce to the entangled SU(2) coherent state (2.19). All examples of coherent states considered in this paper can be constructed accordingly. As an interesting example, we consider the multiple qubit entanglement. Shor introduced the quantum Fourier transforms in order to apply quantum computation to factorize a number $`a`$, with $`0aq.`$ This number $`a`$ can be expressed in qubits as $$|a|\stackrel{}{\epsilon }=|\epsilon _{1,}\text{ }\epsilon _{2,}\text{ }\mathrm{},\epsilon _N,\text{ }\epsilon _i\{0,1\},$$ (A7) where $`\stackrel{}{\epsilon }`$ is the binary notation of $`a.`$ Here $`N`$ is the lowest integer greater than or equal to $`\mathrm{log}_2q.`$ The state $`|a`$ is a product multiparticle SU(2) coherent state for $`j=1/2.`$ The quantum Fourier transform of the state $`|a`$ is $$q^{1/2}\underset{c=0}{\overset{q1}{}}\mathrm{exp}(i2\pi ac/q)|c.$$ (A8) The state $`|c`$ is also a product state for all $`c`$ such that $`0cq1.`$ We can express the transformed state as the entangled SU(2) coherent state (A1) such that $`q^{1/2}{\displaystyle \underset{\stackrel{}{\epsilon }_i\{0,1\}}{}}\mathrm{exp}(i2\pi ac/q)|\stackrel{}{\epsilon }`$ (A9) $`=`$ $`q^{1/2}{\displaystyle \underset{\stackrel{}{\gamma }_i\{0,\mathrm{}\}}{}}\mathrm{exp}(i2\pi ac/q)|j\text{ }\stackrel{}{\gamma }`$ (A10) where $`c\stackrel{}{\epsilon }.`$ The above state can also be written in the form of (A1) with $$f(\stackrel{}{\xi })=q^{1/2}\underset{\stackrel{}{\gamma }_i\{0,\mathrm{}\}}{}\mathrm{exp}(i2\pi ac/q)\delta (\stackrel{}{\xi }\stackrel{}{\gamma }_i).$$ (A11) Hence the Fourier transform state (A5) can be treated within the formalism of entangled SU(2) coherent states.
warning/0001/hep-ph0001221.html
ar5iv
text
# MZ-TH/00–02 January 2000 Δ⁢𝐼=1/2Enhancement and the Glashow-Schnitzer-Weinberg Sum Rule ## Abstract In 1967 Glashow, Schnitzer and Weinberg derived a sum rule in the soft pion and soft kaon limit relating the $`\mathrm{\Delta }I=\frac{1}{2}`$ non-leptonic $`K2\pi `$ amplitude to integrals over strange and non-strange spectral functions. Using the recent ALEPH data from $`\tau `$-decay, we show that the sum rule, slightly modified to reduce contributions near the cut, yields the correct magnitude decay amplitude corresponding to the $`\mathrm{\Delta }I=\frac{1}{2}`$ rule. The $`\mathrm{\Delta }I=\frac{1}{2}`$ rule for kaons has been a challenge to theoreticians for more than four decades, see for a review. The current-current weak non-leptonic Hamiltonian of the Standard Model leads naively to the expectation of roughly equal $`\mathrm{\Delta }I=1/2`$ and $`\mathrm{\Delta }I=3/2`$ amplitudes. Experimentally, however, the $`\mathrm{\Delta }I=1/2`$ amplitudes are enhanced by about a factor 20. QCD corrections may be computed with the help of the operator product expansion yielding a $`\mathrm{\Delta }I=1/2`$ enhancement, but its magnitude turns out too small. The situation may be improved by the use of chiral perturbation theory combined with model input, see for a recent calculation. In this note we will return to the roots and reanalyse an old current algebra calculation of the $`K_s^02\pi `$ matrix element by Glashow, Schnitzer and Weinberg (GSW) . In this classic paper the $`K_s^02\pi `$ amplitude is related to integrals over spectral functions which were in turn evaluated by using the crude approximation of saturation by narrow resonances. As the integrals involve differences of large numbers, it is not astonishing that these authors do not find the right answer. Using the $`K_s^02\pi `$ life time as an imput they end up with a prediction of 8 GeV for the W mass. Since precise data on the relevant spectral functions have recently been obtained by the ALEPH collaboration it seemed prudent to us to reanalyse the GSW formula. It should be pointed out, however, that the ALEPH data do not saturate chiral sum rules even at s as large as $`3GeV^2`$ if substituted directly. In fact, it was shown that if modified chiral sum rules, based on linear combinations of spectral function that vanish at the end of the integration region, are used the saturation is quite spectacular. We will repeat briefly the basic steps of the GSW calculation. In the language of the standard model, the non-leptonic $`\mathrm{\Delta }S=1`$ weak Hamiltonian is given by $$H_W(0)=f^2dxD_{\mu \nu }(x,M_W^2)T\{j_\mu ^{ud}(x)j_\upsilon ^{su}(0)+h.c.\}$$ (1) with $`f^2\frac{G_F}{\sqrt{2}}M_W^2V_{ud}V_{us}^{}`$ where $`M_W`$ is the W boson mass, $`D_{\mu \nu }(x,M_W^2)`$ its propagator, $`G_F`$ the Fermi coupling constant $`(1.166\times 10^5GeV^2)`$, $`V_{ud}`$, $`V_{us}^{}`$ are matrix elements of the CKM matrix and $$j_\mu ^{ud}=\overline{u}\gamma _\mu (1\gamma _5)d,j_\mu ^{su}=\overline{s}\gamma _\mu (1\gamma _5)u.$$ (2) are the strangeness conserving and strangeness changing weak currents, respectively. We consider the decay $`K_s^0\pi ^+\pi ^{}`$ which is described by the matrix element $$𝔐=\pi ^+\pi ^{}\left|H_W\right|K_s^0.$$ (3) Using the conventional ”soft pion” technique and the $`SU(2)\times SU(2)`$ commutation relations yields $$𝔐=\frac{1}{4f_\pi ^2}0\left|H_W\right|K_s^0$$ (4) It should be remembered that the soft pion approach (or $`SU(2)\times SU(2)`$ chiral perturbation theory) gives all non-leptonic K-decay rates in terms of $`𝔐`$, and in accord with the $`\mathrm{\Delta }I=\frac{1}{2}`$ rule. The next step in the derivation is slightly controvertial, namely, GSW evaluate the matrix element $`0\left|H_W\right|K_s^0`$ in the soft kaon limit using the $`SU(3)\times SU(3)`$ algebra of currents. This step is not equivalent to evaluating the original matrix element in the $`SU(3)\times SU(3)`$ chiral limit where the pions and the kaon would have to be treated on the same footing from the beginning and where the amplitude would vanish. Substituting the spectral representation for the vector and axial vector correlators and exchanging the order of integration GSW obtain finally $$𝔐=\frac{3f^2}{64\pi ^2f_\pi ^2f_K}𝑑ss^2\frac{1}{2\pi ^2}\left[(v(s)+a(s))^{ud}(v(s)+a(s))^{us}\right]\left\{\frac{\mathrm{ln}(s/M_W^2)}{1s/M_W^2}\right\}$$ (5) where $`v(s)`$ and $`a(s)`$ are the vector and axial vector spectral functions normalized according to $`v(s)=a(s)=\frac{1}{2}(1+\frac{\alpha _s}{\pi }+\mathrm{})`$ in perurbative QCD. In the derivation of Eq.5 the first and second Weinberg sum rules, which are valid in QCD, have been used. To evaluate the integral in Eq.5, which extends from threshold to $`\mathrm{}`$, we split the integration range into two two parts. From threshold to $`3GeV^2`$ we substitute ALEPH data, and from $`3GeV^2`$ to $`\mathrm{}`$ we use QCD $$\left[(v(s)+a(s))^{ud}(v(s)+a(s))^{us}\right]_{QCD}=\frac{4352}{727}\alpha _s[\alpha _s\overline{u}u(\overline{u}u\overline{s}s)]\frac{1}{s^3}$$ (6) where facrtorization has been used for the four-quark condensate and $`\overline{u}u=\overline{d}d`$ was put. As the QCD part can only serve as an order of magnitude estimate, we should make sure that the integral is saturated by the low energy contribution accessible to experiment. Unfortunately, we know from a detailed study of Weinberg’s and other chiral sum rules that this is not the case. If, however, modified sum rules build up of suitable linear combinations of spectral function sum rules are used which involve integrands that vanish at the end of the given finite integration range, precocious saturation is observed to a surprising extend. This result is understandable from the point of view of global duality in QCD. Analyticity properties of the two-point function relate the integral over a spectral function, multiplied by polynomials, to an integral over a circle in the complex plane. Perurbative QCD is expected to break down close to the time-like real axis, so it is desirable to choose the polynomial so as to vanish at the end of the region of integration over the cut. In the present application we modify the integral Eq.5 by adding a term that vanishes by Weinberg’s sum rule, $`𝔐`$ $`={\displaystyle \frac{3f^2}{64\pi ^2f_\pi ^2f_K}}[{\displaystyle ^R}+{\displaystyle _R^{\mathrm{}}}]dss^2{\displaystyle \frac{1}{2\pi ^2}}\left[(v(s)+a(s))^{ud}(v(s)+a(s))^{us}\right]`$ $`\times \left\{{\displaystyle \frac{\mathrm{ln}(s/M_W^2)}{1s/M_W^2}}{\displaystyle \frac{R}{s}}{\displaystyle \frac{\mathrm{ln}(R/M_W^2)}{1R/M_W^2}}\right\}`$ (7) If the integral is saturated precociously the integral from R to $`\mathrm{}`$ should be negligible. We will assume this to be the case for the time being. In Fig. 1 we plot the result of the first integral over the experimental spectral functions as a function of the upper limit of integration R. As the individual contributions of the non-strange and strange spectral functions are large, we are faced with the difference of two large numbers, and it is the more amazing that the integral appears saturated for $`R1.8GeV^2`$. The small oscillations are typical and will level out for larger R. If the high energy QCD contribution is neglected our prediction is therefore in agreement with the experimental amplitude, $`\left|𝔐\right|_{\mathrm{exp}}=7.78\times 10^7m_K`$ , to within the errors expected from the soft meson approximation. To show that the integral from R to $`\mathrm{}`$ is indeed small we use the estimates $$\alpha _s\overline{u}u^2=1.9\times 10^4GeV^6,\overline{s}s=0.5\times \overline{u}u$$ (8) to find that the high energy QCD contribution is less than 1 % for the lower limit of integration R between 2 to 3 $`GeV^2`$, in a direction bringing the prediction closer to the experimental value. In Fig. 2 we plot the same integral using the original sum rule Eq.5. It is seen that the sum rule converges poorly, but the result is still consistent with the experimental value of the decay amplitude. This is in agreement with the analysis of the non-strange Weinberg sum rules where also only suitable linear combinations of sum rules that vanish at the radius R converge precociously. In Fig.1 and Fig.2 all experimental errors on the spectral functions are suppressed. This is because the statistical part of the errors is washed out completely upon integration and, hopefully, a great part of the (unknown) systematic errors cancel in the differences of spectral functions. We think that this attitude is justified a posteriori by the results plotted in the figures which show the expected shape. We also made an estimate of the small error made by neglecting the contribution of the charm quark. Using the oscillation of the integral with $`sR`$ in Eq.MZ-TH/00–02 January 2000 $`\mathrm{\Delta }I=\frac{1}{2}`$Enhancement and the Glashow-Schnitzer-Weinberg Sum Rule as an estimate of the total error, we arrive at a prediction for the $`K_s^0\pi ^+\pi ^{}`$ matrix element $$\left|𝔐\right|=(9.3\pm .9)\times 10^7$$ (9) as compared to the experimental value $$\left|𝔐\right|_{\mathrm{exp}}=7.78\times 10^7$$ (10) It is remarkable that the simple formula of GSW combined with modern spectral function data should lead to the right prediction of the $`\mathrm{\Delta }I=\frac{1}{2}`$ enhancement. The central value prediction is high by 15 %, i.e. of the order of magnitude expected from errors of the soft kaon approximation and of the neglect of the high energy tail. Had the modern data been available to GSW they would have predicted in 1967 the existence of a charged intermediate vector boson of a mass of about 90 GeV (instead of the 8 GeV following from their assumption of narrow resonance dominance). To quote GSW ” we would lose most of our scruples about this calculation if such an intermediate boson is found”.
warning/0001/nucl-th0001004.html
ar5iv
text
# 1 Introduction ## 1 Introduction The $`(\alpha ,\alpha ^{})`$ reaction on a proton target measured at Saclay has been instrumental in setting the question of the mechanism of $`\mathrm{\Delta }`$ excitation in the projectile (DEP), which was introduced in in order to describe the $`(^3He,t)`$ reaction in p and d targets . That mechanism plays a negligible role in the $`(^3He,t)`$ reaction on proton targets but is quite important in the same reaction on neutron targets and was predicted to be dominant in the $`(^3He,^3He)`$ reaction on proton and neutron targets . Prior to the $`(\alpha ,\alpha ^{})`$ experiment the relevance of the DEP mechanism was a subject of debate , particularly because of the small strength of this mechanism in the $`(^3He,t)`$ reaction on proton targets, which allowed interpretations omitting it . The $`(\alpha ,\alpha ^{})`$ reaction on a proton target is ideal to isolate the DEP mechanism since, because of isospin, the $`\mathrm{\Delta }`$ excitation on the proton target is forbidden. This allowed to test the ideas introduced in Ref and indeed the large peak in the experiment corresponding to DEP was well reproduced. In addition to the issues discussed above, the ($`\alpha ,\alpha ^{})`$ experiment observed a smaller peak at higher excitation energies which was attributed to the Roper excitation. This mode of excitation of the Roper is novel, since it involves an isoscalar source, and can be relevant in determining the strength of three body forces and providing new tools for the comprehension of the $`NNNN\pi \pi `$ and related reactions . The strength of the isoscalar $`NNNN^{}`$ transition was determined empirically from the experimental data in , where the model of for DEP was used and the interference of the two mechanisms was also considered. The analysis proved consistent with the present knowledge of position, decay width and partial decay widths of the Roper and helped to narrow the experimental uncertainties on these magnitudes. The issue of strong isoscalar excitation of nucleon resonances has captured more attention after the first inclusive $`(\stackrel{}{d},d^{})`$ data were obtained . The T<sub>20</sub> data obtained at Dubna on the proton and <sup>12</sup>C, are given in Ref and the final data tables are published in Ref . In Ref polarization observables for the $`(\stackrel{}{d},d^{})`$ reaction on proton targets are discussed bringing new information on electromagnetic form factors of the deuteron and on mechanisms for strong excitation of nucleonic resonances. The description done in Ref was extended to higher energies of around 10-15 GeV , showing that the magnitude of the Roper excitation can be increased by about one order of magnitude and the relative strength of the Roper signal to the one of the DEP mechanism becomes of the order of unity, much bigger than in the $`(\alpha ,\alpha ^{}`$) experiment where it is about 1/4. In Ref polarization observables in the $`(\stackrel{}{p},\stackrel{}{p}^{})`$ on a <sup>4</sup>He target are studied with its view towards possible experiments to be carried out at the Indiana Cyclotron. The program of nucleon resonance excitation using baryonic interactions is thus catching up, and certainly will bring complementary information to the one obtained with electromagnetic probes or meson induced excitation. The present work, DEP and Roper excitation on the $`(d,d^{})`$ reaction on proton targets, should be considered as a complement to the one of the $`(\alpha ,\alpha ^{})`$ reaction . The fact that the deuteron has an isospin $`I=0`$ makes the two works similar since $`\mathrm{\Delta }`$ excitation on the proton target is forbidden in both cases and only DEP and Roper excitation are allowed in the region which we study. However, the fact that the deuteron has a total spin $`J=1`$ induces some differences with respect to the $`(\alpha ,\alpha ^{})`$ reaction and sets different constraints on the theoretical models. With new experiments on this issue coming, the need to have reliable theoretical models to extract the relevant information becomes apparent, and in this sense our present work is a valuable one. We have taken advantage of the existence of experimental information from Saclay measurements of the $`(d,d^{})`$ reaction and we present here a paper where the theoretical ideas are exposed, the data are presented and a discussion is made from comparison of theory and experiment. Early work on the present reaction at different kinematics is done in Ref . We will compare the theoretical results of the model with recent (d,d’) measurements done at Saclay and older (d,d’) measurements done at lower energies and larger angles. ## 2 Formulation In this section we consider a theoretical model of the $`(d,d^{})`$ reaction on the proton target. We include two processes, $`\mathrm{\Delta }`$ excitation in the projectile and Roper excitation in the target, which are shown in Fig. 1 and are the dominant processes in this energy region . We include both the $`\pi N`$ and $`2\pi N`$ decay modes of the Roper resonance. Since we need to take care of the interference between the projectile $`\mathrm{\Delta }`$ process and the target Roper process decaying into $`\pi N`$, we treat the Roper $`\pi N`$ and Roper $`2\pi N`$ processes separately. We take the same model which was used to analyze the $`(\alpha ,\alpha ^{})`$ reaction at $`4.2`$ $`GeV`$ and use the same values for all parameters . The cross section for the $`1\pi `$ decay $`\mathrm{\Delta }`$ and Roper processes is given by, $$\frac{d^2\sigma }{dE_d^{}d\mathrm{\Omega }_d^{}}=\frac{p_d^{}}{(2\pi )^5}\frac{M_d^2M^2}{\lambda ^{1/2}(s,M^2,M_d^2)}d^3p_\pi \frac{1}{E_N^{}\omega _\pi }$$ $$\times \overline{\mathrm{\Sigma }}\mathrm{\Sigma }|T^{1\pi }|^2\delta (E_d+E_NE_d^{}E_N^{}\omega _\pi )$$ (1) where $`\lambda `$ (…) is the Källen function and s the Mandelstam variable for the initial p-d system, and momentum conservation, $`\stackrel{}{p}_d+\stackrel{}{p}_N=\stackrel{}{p}_d^{}+\stackrel{}{p}_N^{}+\stackrel{}{p}_\pi `$ is already implied. The projectile $`\mathrm{\Delta }`$ mechanism ($`T_\mathrm{\Delta }`$) leads to a $`\pi N`$ through the decay of the $`\mathrm{\Delta }`$. Part of the Roper excitation mechanism leads to the same final state through the decay of the $`N^{}`$ into $`\pi N`$. We call this latter piece $`T_{}^{1\pi }`$. Hence the sum of the two mechanisms leading to $`\pi N`$ is given by; $$T^{1\pi }=T_\mathrm{\Delta }+T_{}^{1\pi }$$ (2) The nucleon and deuteron spin sum and average of each $`|T|^2`$ and plus the interference term can be written as, $`\overline{\mathrm{\Sigma }}\mathrm{\Sigma }|T_\mathrm{\Delta }|^2`$ $`=`$ $`{\displaystyle \frac{16}{27}}F_d^2({\displaystyle \frac{f^{}}{\mu }})^4({\displaystyle \frac{f}{\mu }})^2|G_\mathrm{\Delta }|^2({\displaystyle \frac{q^2}{\stackrel{}{q}^2}})`$ (3) $`\times `$ $`[(V_l^{}^2+5V_t^{}^2)\stackrel{}{p}_\mathrm{\Delta }^2+3(V_l^{}^2V_t^{}^2)(\stackrel{}{p}_\mathrm{\Delta }\widehat{q})^2]`$ $$\overline{\mathrm{\Sigma }}\mathrm{\Sigma }|T_{}^{1\pi }|^2=12F_d^2(\frac{f^{}}{\mu })^2g_{\sigma NN}^2g_{\sigma NN^{}}^2|G_{}|^2|D_\sigma F_\sigma ^2|^2\stackrel{}{p}_{}^2$$ (4) $`\overline{\mathrm{\Sigma }}\mathrm{\Sigma }(T_{}^{1\pi }T_\mathrm{\Delta }+T_\mathrm{\Delta }^{}T_{}^{1\pi })`$ (5) $`=`$ $`2Re\{{\displaystyle \frac{16}{3}}F_d^2{\displaystyle \frac{f^{}}{\mu }}({\displaystyle \frac{f^{}}{\mu }})^2{\displaystyle \frac{f}{\mu }}g_{\sigma NN}g_{\sigma NN^{}}D_\sigma ^{}G_{}^{}G_\mathrm{\Delta }F_\sigma ^2`$ $`\times `$ $`[V_l^{}(\stackrel{}{p}_{}\widehat{q})(\stackrel{}{p}_\mathrm{\Delta }\widehat{q})+V_t^{}(\stackrel{}{p}_{}\stackrel{}{p}_\mathrm{\Delta }(\stackrel{}{p}_{}\widehat{q})(\stackrel{}{p}_\mathrm{\Delta }\widehat{q}))]\}\sqrt{{\displaystyle \frac{q^2}{\stackrel{}{q}^2}}}`$ where $`G_\mathrm{\Delta }`$ and $`G_{}`$ are the propagators of the $`\mathrm{\Delta }`$ and Roper resonances, $`D_\sigma `$ the propagator of the $`\sigma `$ meson, $`F_\sigma `$ the $`\sigma NN`$ vertex form factor. The momenta $`\stackrel{}{p}_{},\stackrel{}{p}_\mathrm{\Delta }`$, and $`\stackrel{}{q}`$ are the pion momenta in the Roper rest frame, pion momentum in the $`\mathrm{\Delta }`$ rest frame, and momentum transfer between the nucleons, respectively. $`V_l^{}`$ and $`V_t^{}`$ stand for the longitudinal and transverse parts of $`NNN\mathrm{\Delta }`$ effective interaction which includes $`\pi ,\rho `$ and $`g^{}`$ contributions, where $`g^{}`$ is the Landau-Migdal parameter which is meant to account for short range corrections to the $`\pi `$ and $`\rho `$ exchange. The $`f^{}s`$ and $`g^{}s`$ are coupling constants. All details, including parameter values, are shown in ref. . The function $`F_d`$ is the deuteron form factor defined as $$F_d(\stackrel{}{k})=d^3r\phi ^{}(\stackrel{}{r})e^{i\stackrel{}{k}\frac{\stackrel{}{r}}{2}}\phi (\stackrel{}{r})$$ (6) where $`\phi (r)`$ is the relative wave function of the deuteron obtained from the Bonn potential . The momentum transfer of the deuteron is denoted by $`\stackrel{}{k}=\stackrel{}{p}_d\stackrel{}{p}_d^{}`$ taken in the initial deuteron rest frame. We have included only the s-wave part of the deuteron wave function for simplicity. The contribution from the target Roper process decaying into $`2\pi N`$ is calculated separately as, $$\frac{d^2\sigma }{dE_d^{}d\mathrm{\Omega }_d^{}}=\frac{p_d^{}}{(2\pi )^3}\frac{2M_d^2M}{\lambda ^{1/2}(s,M^2,M_d^2)}\overline{\mathrm{\Sigma }}\mathrm{\Sigma }|T^{\pi \pi }|^2|G_{}|^2\mathrm{\Gamma }_{}^{\pi \pi }$$ (7) with $$\overline{\mathrm{\Sigma }}\mathrm{\Sigma }|T^{\pi \pi }|^2=4F_d^2g_{\sigma NN}^2g_{\sigma NN^{}}^2|D_\sigma F_\sigma ^2|^2$$ (8) using the partial decay width, $`\mathrm{\Gamma }_{}^{\pi \pi }`$, whose explicit form is shown in the Appendix of ref. . This contribution is added to the $`1\pi `$ contributions incoherently. ## 3 Discussion of Experimental results We compare our theoretical results with two independent experimental data sets. One has been obtained recently at $`T_d=2.3GeV`$ at Saturne and another was measured some years ago at $`T_d=1.6GeV`$ . First, we consider the new data which were obtained during a short run at Saturne with a deuteron beam of 2.3 GeV. The deuterons were directed onto a 4 cm thick liquid hydrogen target with thin Ti windows (15 $`\mu m`$). After going through a 40 cm thick lead collimator, the scattered deuterons were momentum analyzed at very small angles using the SPES4 spectrometer with a momentum acceptance of $`\pm `$ 3% and a resolution of $``$ 10<sup>-3</sup>. In the focal plane of the spectrometer, the scattered deuterons were detected using the three front wire chambers of the extended vector polarimeter POMME . In this experiment, special care was taken to minimize all possible experimental backgrounds. A missing mass spectrum was measured at 1.1<sup>o</sup> in five different momentum bites, respectively centered on 2.6, 2.8 3.0, 3.2 and 3.45 GeV/c, covering an excitation energy region up to 600 MeV. The spectrometer acceptance is 15% of the central momentum leading to smaller momentum bites for higher excitation energies. In order to get the most out of the limited beam time in terms of number of counts in each bite and range of coverage in excitation energy, the momentum bites were set as follows: two overlapping bites around 200 MeV of excitation energy, where the excitation of the $`\mathrm{\Delta }`$ resonance in the projectile is expected, and, three non overlapping bites spanning the excitation energy range from 300 to 600 MeV, where the wide Roper resonance is expected (the kinematical limit is at 680 MeV of excitation energy). The target was located outside of the magnetic field of the spectrometer so the usual SPES4 corrections for correlations between scattering angle and scattered momentum were not necessary. The cross section spectrum was binned in 10 MeV steps of excitation energy. For each setting of the spectrometer, empty target measurements were taken for background subtraction. The ratio of full to empty target was in the range of 8 to 10 dropping to 2 for the lowest setting of the spectrometer (corresponding to high excitation energies). We could not get clean measurement at momenta smaller than 2.4 GeV/c because of the large background due to rescattering through the lead collimator. When present, this background shows a strong angular dependence; taking into account the shape of the collimator, different cuts on angular acceptance lead to substantial changes in the shape and slope of the spectrum. For all momentum bites shown in Fig. 2, the off-line analysis of full versus empty target spectra and software cuts on angular acceptances showed no changes in the shape of the spectrum; only overall scaling of the spectrum consistent with the changes of the solid angle, were observed. This extensive off-line analysis led to the conclusion that no significant experimental background was present for these momentum bites. Absolute cross sections were determined using monitors calibrated with the Carbon activation method . This is the well tested standard method of normalization used at Saturne. The measured missing mass spectrum shown in Fig. 2 is dominated by a large structure centered around 200 MeV. This structure has been identified as the excitation of the Delta resonance in the incoming deuteron. The cross section drops sharply between 200 and 300 MeV of excitation energy. Between 300 and 600 MeV, there seems to be an excess of cross section above the high energy tail of the $`\mathrm{\Delta }`$ resonance. There seems to be a discontinuity in the measured cross section around 500 MeV of excitation energy. This corresponds to the highest excitation energy bite that we can cleanly measure and could be affected by the subtraction of a relatively large empty target contribution. In Fig. 2, we compare our theoretical results (including all reaction mechanisms described in section 2) with these experimental data. We have plotted on the same figures the different curves corresponding to the excitation of the $`\mathrm{\Delta }`$ alone, the excitation of the Roper alone and the total cross section taking into account the interference between the two. For the excitation of the Roper, the contributions for the $`\pi `$N and 2$`\pi `$N channels are also shown. We see that the projectile $`\mathrm{\Delta }`$ excitation makes a large contribution to the cross section around 200 MeV. The calculation correctly reproduces the excitation energy of the $`\mathrm{\Delta }`$ and its width, however it over-predicts the cross section at the maximum by at least 20%. The excitation of the $`\mathrm{\Delta }`$ in the projectile cannot account for the observed cross section between 300 and 500 MeV. This range of excitation energy is where we are expecting the Roper resonance to be and the calculation indeed predicts this excess of cross section to be due to the excitation of the Roper resonance. The calculation again overpredicts the measured cross section in this region. In order to extract the excitation of the Roper in the 200 to 600 MeV region, we assume that the shape of the calculated DEP is correct and we normalize it to the experimental spectrum at the maximum of the $`\mathrm{\Delta }`$. This leads to an overall normalization factor of 0.85. We then subtract from the measured data the calculated differential cross section for the excitation of the $`\mathrm{\Delta }`$ and the interference between Roper excitation and $`\mathrm{\Delta }`$ excitation in the projectile, as described in eq. (5), also multiplied by the same normalization factor 0.85. What is left should mainly correspond to the excitation of the Roper resonance plus some physical continuum. In Fig. 3 we plot the measured data points, the predicted calculations normalized by 0.85 and the excess of cross section left once the DEP and interference contributions are subtracted. The excess of cross section has a maximum around 400 MeV, a width at half maximum of about 230 MeV, and an asymmetric shape with a long low energy tail. The calculation of the Roper contribution, normalized by the factor 0.85, agrees qualitatively in shape and strength with the experimental cross section left once we have subtracted the DEP and the interference. Only the theoretical peak is shifted to higher excitation energy by about 25 MeV. The total experimental excess cross section, up to 540 MeV, is 32 $`\pm `$ 8 mb/sr to be compared to the predicted cross section of 28 mb/sr for the Roper resonance. As mentioned earlier, this experimental cross section should be in principle an upper limit since no underlying continuum corresponding to other physical processes has been subtracted. However, in other possible mechanisms were studied, and they were found to be small, leaving only the $`\mathrm{\Delta }`$ excitation in the projectile and Roper excitation in the target as responsible for the reaction cross section in the energy region studied here. According to this, the signal obtained here for the Roper excitation, within experimental and theoretical uncertainties, should be rather fair. An empirical way to subtract the DEP contribution was done in Ref , and this is shown in Fig. 4. The shape of the excitation of the $`\mathrm{\Delta }`$ in the projectile is taken from the measurements of Baldini et al . The assumption being that at T<sub>d</sub>=1.6 GeV, the measured spectrum is mainly dominated by the DEP mechanism and therefore its shape is a good empirical shape for this excitation. This shape is normalized to the data obtained at T<sub>d</sub>=2.3 GeV, at the maximum of the $`\mathrm{\Delta }`$ resonance. Once this empirical DEP contribution and also the interference contribution evaluated as in the previous case are subtracted, a wide structure is left, centered at 350 MeV with a width at half maximum of 230 MeV. The total cross section in this structure (up to 540 MeV of excitation energy) is of the order of 34 $`\pm `$ 8 mb/sr. This value is in agreement with our previous determination of the excess cross section. The shape of the excess cross section is more symmetric than on the previous case and this is possibly due to the fact that the empirical spectrum taken from Ref. contains already some contribution from the excitation of the Roper. We have also compared our theoretical results with data at $`1.6GeV`$ . The data measured at 6.59<sup>o</sup> and 8.05<sup>o</sup> are respectively compared to our predictions in figures 5 and 6. Our calculated results reproduce the overall shape of the spectrum well. However, the theoretical results are about $`30\%`$ smaller than the data at 6.59<sup>o</sup> and about $`40\%`$ at 8.05<sup>o</sup>. This discrepancy has to be looked, however, in the proper perspective. Indeed, the angles where the cross sections are measured in are $`\theta 6.6^0`$. At the smallest angle the cross section has fallen by a factor 30 from the forward direction. This fall down is mostly due to the deuteron form factor which involves large momentum transfers. In this case our neglect of the d-wave in the deuteron is not justified. This, and other approximations could explain these discrepancies. In view of the fact that they represent only about $`1\%`$ of the integrated cross section, we pay no further attention to these discrepancies, but the qualitative agreement found gives also partial support to the model. In what follows we would like to make some estimates of the theoretical uncertainties in the present analysis of the data. One of the sources of uncertainty is our neglect of the d-wave in the deuteron wave function. The other one is the possible effect of Fermi motion in the deuteron . These two factors could change the shape on the $`\mathrm{\Delta }`$ excitation strength and hence lead to uncertainties in the Roper excitation after the DEP strength and interference are subtracted from the data. We begin by the effect of Fermi motion. The deuteron Fermi motion is considered here in the same way as done in . It affects the $`\mathrm{\Delta }`$ propagator which enters the evaluation of the DEP mechanism. This propagator is given by $$G_\mathrm{\Delta }(s)=\frac{1}{\sqrt{s}M_\mathrm{\Delta }+\frac{i}{2}\mathrm{\Gamma }_\mathrm{\Delta }(s)}$$ (9) where the variable s is taken as $$s=(q^0+M)^2\left(\frac{\stackrel{}{q}+\stackrel{}{p}_\pi }{2}\right)^2,$$ (10) where $`q`$ and $`p_\pi `$ are the momenta of the exchanged meson, and the emitted pion, respectively , taken in the frame of reference where the deuteron is at rest. In this approximation the momentum transfer is shared equally by the initial and final nucleon in the deuteron. The fairness of this approximation to account for Fermi motion of the nucleus was well established in Refs in the study of coherent pion photoproduction with similar momentum transfers as here. However, in order to see the effects of Fermi motion and have a feeling for possible uncertainties from this source we have conducted new calculations in which in eq. (10) we assume the initial momentum of the struck nucleon of the deuteron to be zero. This replaces $`\frac{\stackrel{}{q}+\stackrel{}{p_\pi }}{2}`$ in that equation by $`\stackrel{}{q}`$. We can see the results of the new calculation in Fig.7. We can see that the strength of the $`\mathrm{\Delta }`$ excitation is increased by about 20 %. The prescription followed here to account for Fermi motion was found in Refs to be rather accurate, but even then we see that the effects of ignoring it altogether does not bring drastic changes in the cross section. The possible uncertainties from this source are further minimized if we normalize the theoretical results to the experimental cross section, as we have done in the analysis of this work. Indeed, if we do so we obtain the results shown in Fig. 8, which are shown superposed to those of Fig.3. As we can see there, the differences found between neglecting the Fermi motion , or taking it according to our prescription, are very small up to 500 MeV of excitation energy, once the normalization of the cross section around the peak of the delta resonance is done. As for the d-wave of the deuteron we have proceeded as follows: In the evaluation of the deuteron form factor of eq. (6) we have taken only the s-wave of the deuteron so far. Inclusion of the d-wave into our scheme would lead to two parts. One with the same structure as we have, which goes with the $`j_0(\frac{kr}{2})`$ component of the exponential, but substituting $`u^2`$ by $`u^2+w^2`$ in eq. (6) ($`u`$ and $`w`$ are the s and d-wave parts of the deuteron wave function, respectively), and another one which goes with the $`j_2(\frac{kr}{2})`$ component of the exponential. Detailed evaluations of these two parts in the deuteron form factor can be seen in and there we see that up to 500 MeV/c the $`j_2`$ part of the form factor contributes less than 10%. On the other hand the difference between the $`j_0`$ component evaluated with just the s-wave (normalized to unity) on the s plus d-wave parts of the deuteron wave function are smaller than 4% up to this momentum. In our case 500 MeV/c momentum transfer corresponds to an excitation energy of around 550 MeV in Fig. 2, just the tail of the distribution beyond the Roper excitation region which we have studied here. On the other hand in the case of Fig. 5 , 500 MeV/c would appear at $`p_d`$ around 2.25 GeV/c and in Fig. 6 one has already 500 MeV/c momentum transfer around the peak of the distribution. Hence, our neglect of the d-wave part of the wave function would induce more uncertainties, in the line we discussed above when we discussed the Baldini’s data. Altogether, we can safely say that in our analysis of the Roper excitation the uncertainties coming from the theoretical model and approximations done are at the level of 10-15%. ## 4 Conclusion With the help of a theoretical model previously used to analyze the $`(\alpha ,\alpha ^{})`$ reaction on the proton, exciting the $`\mathrm{\Delta }`$ in the projectile plus the Roper, we have analyzed data on the $`(d,d^{})`$ reaction on proton targets at a deuteron energy 2.3 GeV. The use of the model becomes necessary because there is an important interference between the mechanism of delta excitation in the projectile and Roper excitation in the target (followed by $`\pi N`$ decay). We observed that the model gave a good reproduction of the shape of the $`\mathrm{\Delta }`$ excitation in the target, but the normalization exceeded the data by about $`20\%`$. In view of that in order to subtract this contribution and obtain the strength for Roper excitation, we found it justified to normalize the theoretical results by a factor 0.85 which leads to good agreement with the data in the $`\mathrm{\Delta }`$ excitation region. Similarly we multiplied by the same factor the interference term, calculated theoretically, and these two pieces of ”background” were subtracted from the measured data in order to obtain the Roper excitation strength. We found a qualitative agreement with the theoretical predictions for Roper excitation of the model, also normalized by the same factor. In order to estimate uncertainties due to the use of a theoretical model in the analysis we compared our results to those obtained in Ref , where an empirical approach was used to subtract the DEP contribution using the shape of the $`\mathrm{\Delta }`$ resonance excitation from a previous measurement at the lower energies, where there should be a small contribution of Roper excitation. The results obtained with both methods agree qualitatively, the integrated strengths obtained are similar, only the peaks of the Roper strength appear a bit shifted with respect to each other in the two analyses. We estimate that considering statistical and systematic errors, the latter ones from the model dependence of the subtractions, the strength of the Roper determined here is accurate within $`25\%`$, and within these errors the agreement with theory can be claimed acceptable. The shape and the width of the Roper strength are compatible with the empirical information about the resonance. The present analysis confirms the substantial strength for $`NNNN^{}`$ transition in the scalar channel which has been shown already to have important repercussions in different physical phenomena at intermediate energies. These results, and the importance that the isoscalar excitation of the Roper is bound to have in other process, should stimulate further experiments at higher energies. ## 5 Acknowledgments We are grateful to the COE Professorship of Monbusho which enabled one of us, E. O., to stay at RCNP where part of this work has been done. One of us, S. H., acknowledges the hospitality of SATURNE, Saclay during his stay and the hospitality of the University of Valencia where part of this work was done. This work is partly supported by DGICYT contract no. PB 96-07053 and an NSF grant. We are also grateful to Dr E. Tomasi-Gustafsson for very helpful discussions. Figure Caption Fig. 1 Diagrams for the p(d,d’)X reactions considered in this paper. They are (a) the $`\mathrm{\Delta }`$ excitation in the deuteron and (b) the Roper excitation in the proton. The $`\sigma `$ exchange must be interpreted as an effective interaction in the isoscalar exchange channel . Fig. 2 The double differential cross section $`d^2\sigma /dM_Id\mathrm{\Omega }`$ is shown as a function of excitation energy of the proton. The $`M_I`$ is the invariant mass of the target system. Solid circles indicate the experimental data obtained in ref. . The theoretical calculations are also shown in the figure, which are total spectrum (solid line), contribution from $`\mathrm{\Delta }`$ excitation (dashed line), and contribution from Roper excitation (thick dotted line). The Roper contributions decaying into $`\pi N`$ and $`\pi \pi N`$ are separately shown as thin dotted lines. Fig. 3 The double differential cross section $`d^2\sigma /dM_Id\mathrm{\Omega }`$ is shown as a function of excitation energy of the proton. The $`M_I`$ is the invariant mass of the target system. Solid circles indicate the experimental data obtained in ref. . The theoretical calculations normalized by factor 0.85 are also shown in the figure, which are total spectrum (solid line), contribution from $`\mathrm{\Delta }`$ excitation (dashed line), and contribution from Roper excitation (dotted line). Solid triangles are Roper contribution extracted from the data by subtracting the calculated $`\mathrm{\Delta }`$ and interference contributions with the normalization factor 0.85. Fig. 4 The double differential cross section $`d^2\sigma /dM_Id\mathrm{\Omega }`$ is shown as a function of excitation energy of the proton. The $`M_I`$ is the invariant mass of the target system. Solid circles indicate the experimental data obtained in ref. . The solid line indicates the $`\mathrm{\Delta }`$ excitation contribution extracted from data obtained in Ref. and normalized to present data at the peak. The Roper contribution calculated by our model and normalized by the factor 0.85 is shown by dotted line. Solid triangles are Roper contribution extracted from the data by subtracting the $`\mathrm{\Delta }`$ contribution shown by the solid line and the calculated interference contribution with the normalization factor 0.85. Fig. 5 The double differential cross section $`d^2\sigma /dpd\mathrm{\Omega }`$ is shown as a function of the emitted deuteron momentum at $`T_d`$=1.6 GeV. Solid circles indicate the experimental data obtained in Ref. . The theoretical calculations are also shown in the figure, which are total spectrum (solid line), contribution from $`\mathrm{\Delta }`$ excitation (dashed line), and contribution from Roper excitation (dotted line). Fig. 6 The double differential cross section $`d^2\sigma /dpd\mathrm{\Omega }`$ is shown as a function of the emitted deuteron momentum at $`T_d`$=1.6 GeV. Solid circles indicate the experimental data obtained in Ref. . The theoretical calculations are also shown in the figure, which are total spectrum (solid line), contribution from $`\mathrm{\Delta }`$ excitation (dashed line), and contribution from Roper excitation (dotted line). Fig. 7 The double differential cross section $`d^2\sigma /dM_Id\mathrm{\Omega }`$ is shown as a function of excitation energy of the proton. The $`M_I`$ is the invariant mass of the target system. Calculated total spectrum (solid line) and $`\mathrm{\Delta }`$ excitation contribution (dashed line) are shown. Thin lines are the same results as shown in Fig. 2. Thick lines indicate the results obtained neglecting the Fermi motion of nucleon in the projectile, see text. Fig. 8 Same as Fig. 7 except for the normalization to the experimental strength. Thin lines correspond to the calculated results in Fig. 3.
warning/0001/quant-ph0001033.html
ar5iv
text
# Adiabatic Output Coupling of a Bose Gas at Finite Temperatures ## I Introduction Trapped atomic Bose-Einstein Condensates (BEC) are now routinely produced in various laboratories around the world, and it is important to understand the factors that influence the coherence of atoms transferred from them. This is an essential issue for the atom laser research, which has the long-term goal of producing continuous, directional, and coherent beams of atoms. A matter-wave pulse was first produced by using a radio frequency (RF) electromagnetic pulse to transfer atoms out of a trap, where they were allowed to fall freely under gravity. More recently, a stimulated Raman process induced a transition to an untrapped magnetic state in an experiment at NIST; net momentum kick provided by the process resulted in a highly directional beam. On the other hand, a long beam of atoms falling under gravity was produced in Munich using an RF-field-induced transition. It is noted that although the more general features of the output in these experiments are fairly well-understood, detailed properties of the atoms in the output beam, and the evolution of the component that remains inside the trap have not so far been investigated. Previous theoretical treatments of the output couplers for condensates have been either limited to a single-mode non-interacting trapped condensate or to mean field treatment for the condensate which assume that the output beam is extracted out of a condensate at zero temperature and that they can be described by a single complex function of space and time. However, real condensates appear at finite temperatures, and as a result, thermal excitations play a major role. In a previous paper, we outlined a theory of weak output coupling from a partially condensed, trapped Bose gas at finite temperatures. By applying the self-consistent Hartree-Fock-Bogoliubov (HFB) theory for Bose gases at finite temperatures, we identified three output components. The first is the output of pure condensates, which we have called “coherent output.” The second is the fraction emerging from the thermal excitations in the trap, which are coupled out of the trap by the process of “stimulated quantum evaporation.” This is equivalent to the quantum evaporation of Helium atoms from the surface of superfluid <sup>4</sup>He, where phonon excitations travel up to the surface of the superfluid and then spontaneously emerge from the surface as evaporated atoms. In our case such an evaporation from the trap is stimulated by an electromagnetic field. The last output component comes from the process of “pair breaking,” which involves simultaneous creation of an output coupled atom and an elementary excitation (quasi-particle) within the trap. For suitable choices of the coupling parameters each of the three processes can become the dominant process. We have shown that output coupling can serve not only as a useful way to extract an atomic beam out of a trap, but also as a probe to the delicate features of the quantum state of the Bose gas, including the pair correlations inside the condensate. In this paper, we present an extensive analysis of the spectrum of the output atoms, and address issues that were not included in our shorter work. The first is the conditions for the output coupling to give a steady flow of atoms. We discuss the behaviour of the output rate and the atomic density in the short and the long time regimes, and also discuss the conditions for achieving a steady output beam. Second, the application of output coupling must cause changes in the state of the trapped Bose gas, such as changes of the number of excitations relative to the number of condensate atoms in the trap. We present a thorough discussion of these changes. The state of the Bose gas in the trap is usually described by the Bogoliubov formalism, which assumes an indefinite number of atoms in the system and therefore is not number-conserving. In this paper we discuss a number-conserving description of the system, which is especially useful when we consider the process of pair-breaking. The results of this paper are directly applicable for any output coupling scheme which involves a single trapped state and one output state. We demonstrate here general fundamental issues by considering a one-dimensional Bose gas in a harmonic potential, which is coupled into a free output level in the absence of gravity. The structure of this paper is as follows: We begin by deriving the equations of motion for the evolution of dynamical variables inside and outside the trap in Section II. We present in Section III a quasi-steady-state formalism, which enables us to obtain the properties of the output atoms. We demonstrate the results by a numerical example. We outline in Section IV the solution to the equations of motion in the trap that were derived in Section II, by introducing a number-conserving, time dependent Hartree-Fock-Bogoliubov (HFB) formulation in an adiabatic approximation. Applying this, we obtain expressions for the internal modes of the system in two different regimes, from which the time dependent quasiparticle excitations can be calculated. Finally, discussions and summary are given in Section V. ## II The two-state output coupling model In this section we present our model for describing the output coupling of a trapped Bose gas into free output modes. We derive the equations of motion for the atomic field operators in the trapped and the untrapped states, and give a general form of their solutions. ### A Description of the model Our model assumes that atoms are initially in an atomic magnetic level $`|t`$ (“the trapped state”) confined by a potential and in thermal equilibrium. A coupling interaction is then switched on, inducing transitions to a different magnetic level $`|f`$ (the “free” or “untrapped” state). We stress that these are labels denoting internal atomic levels, not the centre of mass states, so that for short enough times an atom in an $`|f`$ state may still be present within the trap. We use the atomic field operator $`\widehat{\psi }_t(𝐫)`$ to describe the amplitude for the annihilation of a trapped atom at point $`𝐫`$, and the operator $`\widehat{\psi }_f(𝐫)`$ to describe the corresponding amplitude for a free, untrapped atom. The Hamiltonian of the system takes the form $$\widehat{}=\widehat{}_t^{(0)}+\widehat{}_f^{(0)}+\widehat{}_{\mathrm{couple}},$$ (1) where $`\widehat{}_t^{(0)}`$ and $`\widehat{}_f^{(0)}`$ describe the dynamics of the trapped and untrapped atoms respectively while $`\widehat{}_{\mathrm{couple}}`$ describes the coupling between the two states. The dynamics inside the trap are given by the many-body Hamiltonian: $`\widehat{}_t^{(0)}`$ $`=`$ $`{\displaystyle d^3𝐫\widehat{\psi }_t^{}(𝐫)\left[\frac{\mathrm{}^2^2}{2m}+V_t(𝐫)\right]\widehat{\psi }_t(𝐫)}`$ (3) $`+{\displaystyle \frac{1}{2}}{\displaystyle d^3𝐫d^3𝐫^{}\widehat{\psi }_t^{}(𝐫)\widehat{\psi }_t^{}(𝐫^{})U_{tt}(𝐫𝐫^{})\widehat{\psi }_t(𝐫^{})\widehat{\psi }_t(𝐫)},`$ where $`m`$ is the mass of a single atom, $`V_t(𝐫)`$ is the potential responsible for the confinement of the atoms in the trap and $`U_{tt}`$ is the inter-particle potential between the trapped atoms. With the output atoms, significant effect of their (elastic) collisions with the trapped atoms must be taken into account. In addition, we consider a small rate of output from the trap, so that the output atoms are dilute; this enables us to neglect the interactions between the free atoms themselves. Since the Bose gases are typically so dilute that their mean-free-path for the inelastic collisions is much larger than the dimensions of the atomic cloud in the trap, one may neglect also any inelastic collisions of the output atoms with the trapped atoms. The Hamiltonian for the output atoms is then given by $`\widehat{}_f^{(0)}`$ $`=`$ $`{\displaystyle d^3𝐫\widehat{\psi }_f^{}(𝐫)\left[\frac{\mathrm{}^2^2}{2m}+V_f(𝐫)\right]\widehat{\psi }_f(𝐫)}`$ (5) $`+{\displaystyle d^3𝐫d^3𝐫^{}\widehat{\psi }_f^{}(𝐫)\widehat{\psi }_t^{}(𝐫^{})U_{tf}(𝐫𝐫^{})\widehat{\psi }_f(𝐫^{})\widehat{\psi }_t(𝐫)},`$ where $`V_f(𝐫)`$ is the potential that influences the propagation of the output free atoms and $`U_{tf}`$ is the collisional interaction between the trapped and free atoms; this is in general different from the interaction $`U_{tt}`$. We use the usual $`\delta `$-function form for the inter-particle potentials $`U_{tt}(𝐫𝐫^{})`$ $`=`$ $`U_0\delta (𝐫𝐫^{})`$ (6) $`U_{tf}(𝐫𝐫^{})`$ $`=`$ $`U_1\delta (𝐫𝐫^{}),`$ (7) where $`U_0=4\pi \mathrm{}^2a_{tt}/m`$ and $`U_1=4\pi \mathrm{}^2a_{tf}/m`$ are proportional to the $`s`$-wave scattering lengths $`a_{tt}`$ and $`a_{tf}`$ for trapped-trapped and trapped-free collisions, respectively. We assume a repulsive interaction between the atoms, i.e. $`U_{0,1}>0`$. For the Hamiltonian $`\widehat{}_{\mathrm{couple}}`$, we consider coupling by an electromagnetic (EM) field which induces transitions between the states $`|t`$ and $`|f`$. In the rotating wave approximation, the EM coupling mechanism may be described by the following Hamiltonian, which can quite clearly be generalised to describe any kind of linear coupling such as weak tunnelling: $$\widehat{}_{\mathrm{couple}}=\mathrm{}d^3𝐫\lambda (𝐫,t)\widehat{\psi }_f^{}(𝐫)\widehat{\psi }_t(𝐫)+h.c.$$ (8) Here $`\lambda (𝐫,t)`$ denotes the amplitude of coupling between the trapped and the untrapped magnetic states. The form of $`\lambda (𝐫,t)`$ depends on the type of coupling used: Typical mechanisms are direct (one-photon) radio-frequency transition and indirect (two-photon) stimulated Raman transition. For any EM induced processes, the coupling can be written as $$\lambda (𝐫,t)=\overline{\lambda }(𝐫,t)e^{i(𝐤_{\mathrm{em}}𝐫\mathrm{\Delta }_{em}t)},$$ (9) where $`\overline{\lambda }`$ is slowly varying in space and time. $`\overline{\lambda }`$ can be either time-independent, to describe a continuous electromagnetic wave, or pulsed. $`\mathrm{}𝐤_{\mathrm{em}}`$ and $`\mathrm{}\mathrm{\Delta }_{\mathrm{em}}`$ measure the net momentum and energy transfer from the EM field to an output atom. In an RF coupling scheme, $`\overline{\lambda }(𝐫,t)`$ is the Rabi frequency $`\mathrm{\Omega }(𝐫,t)=\widehat{p}(𝐫,t)/\mathrm{}`$ (or $`\widehat{\mu }(𝐫,t)/\mathrm{}`$) corresponding to the flipping of the atomic electric (or magnetic) dipole $`\widehat{p}`$ (or $`\widehat{\mu }`$) in the electric (or magnetic) field $`(𝐫,t)`$ ($`(𝐫,t)`$); $`\mathrm{\Delta }_{\mathrm{em}}`$ is the detuning of the EM field frequency from the transition frequency and $`𝐤_{\mathrm{em}}`$ is, in general, negligible compared to the initial momentum distribution of the atoms. In the stimulated Raman coupling, two laser beams are used to induce a transition from $`|t`$ to $`|f`$ through an intermediate level $`|i`$, and $$\overline{\lambda }(𝐫,t)=\frac{\mathrm{\Omega }_{ti}^{}(𝐫,t)\mathrm{\Omega }_{fi}(𝐫,t)}{\mathrm{\Delta }_i},$$ (10) where $`\mathrm{\Omega }_{ti}`$ and $`\mathrm{\Omega }_{fi}`$ are the Rabi frequencies corresponding to the intermediate transitions and $`\mathrm{\Delta }_i`$ is their detuning from resonance with the two beams. $`\mathrm{\Delta }_{\mathrm{em}}`$ and $`𝐤_{\mathrm{em}}`$ are the differences between the frequencies and momenta associated with the two laser beams: $`\mathrm{\Delta }_{em}`$ $`=`$ $`\omega _{1L}\omega _{2L}{\displaystyle \frac{E_t^{(0)}E_f^{(0)}}{\mathrm{}}},`$ (11) $`𝐤_{em}`$ $`=`$ $`𝐤_{1L}𝐤_{2L},`$ (12) where $`E_t^{(0)}E_f^{(0)}`$ is the energy splitting between the atomic levels $`|t`$ and $`|f`$ in the centre of the trap. A more detailed derivation of Eq. (10) for the Raman process is provided in Appendix A. An energy level diagram depicting the output coupling through the stimulated Raman process is given in Fig. 1. ### B Equations of motion The coupled equations of motion for the trapped and free field operators are obtained by computing their commutation relations with the Hamiltonian (1). We thus find $`{\displaystyle \frac{}{t}}\widehat{\psi }_t(𝐫,t)`$ $`=`$ $`{\displaystyle \frac{i}{\mathrm{}}}_t\widehat{\psi }_t(𝐫,t){\displaystyle \frac{i}{\mathrm{}}}U_0\widehat{\psi }_t^{}(𝐫,t)\widehat{\psi }_t(𝐫,t)\widehat{\psi }_t(𝐫,t)`$ (14) $`i\lambda ^{}(𝐫,t)\widehat{\psi }_f(𝐫,t),`$ $`{\displaystyle \frac{}{t}}\widehat{\psi }_f(𝐫,t)`$ $`=`$ $`{\displaystyle \frac{i}{\mathrm{}}}_f\widehat{\psi }_f(𝐫,t)i\lambda (𝐫,t)\widehat{\psi }_t(𝐫,t),`$ (15) where $`_t`$ $``$ $`\mathrm{}^2^2/2m+V_t(𝐫),`$ (16) $`_f`$ $``$ $`\mathrm{}^2^2/2m+V_f(𝐫)+U_1\widehat{\psi }_t^{}(𝐫)\widehat{\psi }_t(𝐫).`$ (17) In Eq. (17) we have used a mean-field approximation for the collisional effect of the trapped atoms on the untrapped ones. This approximation neglects inelastic scattering processes with the trapped atoms and other possible effects of entanglement of the output atoms with the atoms in the trap. The approximation is justified under the assumption of long atomic mean-free-path mentioned above. ### C General solutions #### 1 Output atoms The formal solution of Eq. (15) for $`\widehat{\psi }_f`$ in terms of $`\widehat{\psi }_t`$ is $$\widehat{\psi }_f(𝐫,t)=\widehat{\psi }_f^{(0)}(𝐫,t)i_0^t𝑑t^{}d^3𝐫^{}K_f(𝐫,𝐫^{},tt^{})\lambda (𝐫^{},t^{})\widehat{\psi }_t(𝐫^{},t^{}),$$ (18) where $`\widehat{\psi }_f^{(0)}`$ satisfies the time-dependent Schrödinger equation for the free evolution of $`\widehat{\psi }_f`$ in the absence of output coupling $$\frac{}{t}\widehat{\psi }_f^{(0)}=\frac{i}{\mathrm{}}_f\widehat{\psi }_f^{(0)},$$ (19) and the “free” propagator $`K_f(𝐫,𝐫^{},tt^{})`$ satisfies the partial differential equation $$\frac{}{t}K_f(𝐫,𝐫^{},tt^{})=\frac{i}{\mathrm{}}_fK_f(𝐫,𝐫^{},tt^{})+\delta (𝐫𝐫^{})\delta (tt^{}).$$ (20) The second term in Eq. (18) describes the transition of the atoms from the trapped level $`|t`$ into the free level $`|f`$ with amplitude $`\lambda (𝐫,t)`$ and their subsequent propagation as free atoms. We note that it is useful to expand the field operator $`\widehat{\psi }_f`$ in terms of the normal modes $`\phi _𝐤(𝐫)`$ of the untrapped level $$\widehat{\psi }_f(𝐫,t)=\underset{𝐤}{}\phi _𝐤(𝐫)\widehat{b}_𝐤(t),$$ (21) where $`\widehat{b}_𝐤`$ satisfy the usual bosonic commutation relations $`[\widehat{b}_𝐤,\widehat{b}_𝐤^{}^{}]=\delta _{𝐤,𝐤^{}}`$. $`𝐤`$ denotes the momentum state of the free atoms with energy $`E_𝐤=\mathrm{}\omega _𝐤`$, and $`\phi _𝐤`$ are the time-independent solutions of the single-particle problem in the non-trapping effective potential $`V_f(𝐫)+U_1\widehat{\psi }_t^{}(𝐫)\widehat{\psi }_t(𝐫)`$ created by the mean field effect of the collisions between the trapped and the free atoms. In principle, the solutions $`\phi _𝐤`$ and the energies $`E_𝐤`$ may change with time due to the change in the density of the trapped atoms and the subsequent change in the effective repulsive potential near the trap. In what follows we will neglect this time-dependence under the assumption of weak output coupling and slow changes in the density of the trapped atoms. In the absence of gravitational or other forces, at positions far away from the trap, $`𝐤`$ may be taken to be the wave number of a plane wave $`\phi _𝐤e^{i𝐤𝐫}`$ with $`\omega _𝐤=\mathrm{}k^2/2m`$. In the presence of gravity, the structure of the output modes should be defined appropriately, and the modes $`𝐤`$ may be given asymptotically by the solutions of the Schrödinger equation in a homogeneous field. In Eq. (21) we have used a sum $`_𝐤`$ over discrete output states. It is noted that the actual structure of the Hilbert space for the output modes depends on the potential $`V_f(𝐫)`$; if this potential vanishes far away from the centre of the trap, then the sum $`_𝐤`$ should be replaced by an integral $`d^3𝐤`$ and the operators $`\widehat{b}_𝐤`$ should be defined such that $`[\widehat{b}_𝐤,\widehat{b}_𝐤^{}^{}]=\delta (𝐤𝐤^{})`$. In terms of the basis functions $`\phi _𝐤(𝐫)`$, the free field operator $`\widehat{\psi }_f^{(0)}`$ is given by $$\widehat{\psi }_f^{(0)}(𝐫,t)=\underset{𝐤}{}\phi _𝐤(𝐫)\widehat{b}_𝐤(0)e^{i\omega _𝐤t}$$ (22) and the propagator of the free atoms may be written as $$K_f(𝐫,𝐫^{},tt^{})=\underset{𝐤}{}\phi _𝐤(𝐫)\phi _𝐤^{}(𝐫^{})e^{i\omega _𝐤(tt^{})}\theta (tt^{}).$$ (23) It is useful to describe the evolution of the untrapped atoms in terms of the annihilation operators $`\widehat{b}_𝐤`$ in a specific mode $`𝐤`$. The solution for this operator is obtained by multiplying Eq. (18) by $`\phi _𝐤^{}`$ and integrating over all space: $$\widehat{b}_𝐤(t)=\widehat{b}_𝐤(0)e^{i\omega _𝐤t}i_0^t𝑑t^{}d^3𝐫\phi _𝐤^{}(𝐫)\lambda (𝐫,t^{})e^{i\omega _𝐤(tt^{})}\widehat{\psi }_t(𝐫,t^{}).$$ (24) Solutions for the output field $`\widehat{\psi }_t(𝐫,t)`$ in Eqs. (18) and (24) require an explicit expression for the trapped field operator $`\widehat{\psi }_t(𝐫,t^{})`$. The simplest approximation is to take the first order solution in the coupling amplitude $`\lambda (𝐫,t)`$. This corresponds to a very weak coupling and $`\widehat{\psi }_t(𝐫,t)\widehat{\psi }_t^{(0)}(𝐫,t)`$, where $`\widehat{\psi }_t^{(0)}(𝐫,t)`$ is the field operator of the trapped atoms without output coupling. The fundamental properties of the output under such approximation will be the main subject of Sec. III. #### 2 Trapped atoms By substituting the solution (18) for $`\widehat{\psi }_f`$ back into Eq. (14) we obtain the following equation for the trapped field operator $`\widehat{\psi }_t(𝐫,t)`$ $`{\displaystyle \frac{}{t}}\widehat{\psi }_t(𝐫,t)`$ $`=`$ $`{\displaystyle \frac{i}{\mathrm{}}}_t\widehat{\psi }_t(𝐫,t){\displaystyle \frac{i}{\mathrm{}}}U_0\widehat{\psi }_t^{}(𝐫,t)\widehat{\psi }_t(𝐫,t)\widehat{\psi }_t(𝐫,t)`$ (26) $`{\displaystyle _0^t}𝑑t^{}{\displaystyle d^3𝐫G(𝐫,𝐫^{},t,t^{})\psi _t(𝐫^{},t^{})}i\lambda ^{}(𝐫,t)\widehat{\psi }_f^{(0)}(𝐫,t),`$ where $$G(𝐫,𝐫^{},t,t^{})=\lambda ^{}(𝐫,t)K_f(𝐫,𝐫^{},tt^{})\lambda (𝐫^{},t^{}).$$ (27) The solution of the integro-differential equation, Eq. (26) will be the main subject of Sec. IV. In principle, two different situations may be expected from such an integro-differential equation. In the case where this equation describes a coupling to the output levels with a narrow available bandwidth compared to the coupling strength $`\lambda (𝐫,t)`$, we anticipate Rabi oscillations of the atomic population between the trapped and untrapped levels. Physically, this means that the output atoms stay near the trap for a long enough time to perform these oscillations. On the other hand, when the bandwidth of the output modes is large compared to the coupling strength, an exponential decay of the population in the trap is expected. Physically, this behaviour is expected when the output atoms are fast enough to escape from the trap before the interaction couples them back into the trapped level. Even if the coupling is very weak, an oscillatory kind of behaviour is expected for short times compared to the inverse of the bandwidth of the relevant output modes, before the output rate settles on a constant rate with fixed energy. For this last case, Eq. (26) may be viewed as a Langevin equation for an interaction of a confined system with an infinite heat bath. The trapped field operators $`\widehat{\psi }_t(𝐫,t)`$ can, in principle, be expanded similarly to the field operator $`\widehat{\psi }_f(𝐫,t)`$ as $$\widehat{\psi }_t(𝐫,t)=\underset{n}{}\varphi _n(𝐫)\widehat{a}_n(t),$$ (28) where $`\varphi _n(𝐫)`$ are the normal modes of the trap given by the solutions of the single-particle problem in the potential well $`V_t(𝐫)`$. These eigenmodes, however, do not form a good basis, since the interaction between the atoms results in a strong mixing between the levels. In the following sections, therefore, we will instead use the basis of condensate and its quasi-particle excitations, which are obtained from the Hartree-Fock-Bogoliubov theory of an interacting Bose gas. ## III Output properties in the quasi-steady state In this section we present the properties of the output atoms under the assumption that the output coupling is very weak. In this case, the output beam of atoms can serve as a probe to the structure of the Bose gas in the trap under steady-state conditions. The solution for the output is given by Eq. (18), or equivalently Eq. (24), where we substitute $`\widehat{\psi }_t(𝐫,t)\widehat{\psi }_t^{(0)}(𝐫,t)`$. We start this section by giving a brief description of the formalism that allows the use of the steady-state solution $`\widehat{\psi }_t^{(0)}(𝐫,t)`$ for the atoms in the trap. We then present basic properties of the output, in particular the spectrum and the density of the output coupled atoms. Finally the first and second order correlations of the output atoms are presented. ### A The trapped atoms in the quasi steady-state We briefly review in this subsection the theory of the trapped Bose gas in steady-state conditions, in a way that will enable us in Sec. IV to extend the theory to the time-dependent case where the number of atoms in the trap does change adiabatically during output coupling. Since the situations discussed in this paper involve transitions of atoms into untrapped propagating states where counting the number of output atoms could be one of the main measurements, we choose to use a number-conserving theory in the spirit of the theories put forward recently. In addition, the theory enables an extension of the finite temperature HFB-Popov method into time dependent cases. For describing a partially condensed system of atoms at a finite temperature, we split the field operator for the atoms in the trap into a part which is proportional to the condensate wave function and a part which represents excitations orthogonal to this state: $$\widehat{\psi }_t(𝐫,t)=e^{i\mathrm{\Phi }(t)}\left[\psi _0(𝐫)\widehat{a}_0(t)+\delta \widehat{\psi }(𝐫,t)\right].$$ (29) Here $`\mathrm{\Phi }(t)`$ is a global phase given by $$\mathrm{\Phi }(t)=_0^t\mu (t^{})𝑑t^{},$$ (30) where $`\mu `$, the chemical potential of the system for the given global variables, is constant under steady-state conditions. The operator $`\widehat{a}_0`$ is a bosonic annihilation operator satisfying $`[\widehat{a}_0,\widehat{a}_0^{}]=1`$ and describes the annihilation of one atom in the condensate state $`\psi _0(𝐫)`$. The number of condensate atoms is represented by the operator $`\widehat{N}_0\widehat{a}_0^{}\widehat{a}_0`$. The non-condensate part, $`\delta \widehat{\psi }`$, is assumed to be orthogonal to the condensate in the sense $`d^3𝐫\widehat{\psi }_0^{}(𝐫,t)\delta \widehat{\psi }(𝐫,t)=0`$. In the number conserving formalism $`\delta \widehat{\psi }`$ is approximated by the following Bogoliubov form: $$\delta \widehat{\psi }(𝐫,t)\widehat{a}_0\frac{1}{\sqrt{\widehat{N}_0}}\underset{j}{}[u_j(𝐫)\widehat{\alpha }_j(t)+v_j^{}(𝐫)\widehat{\alpha }_j^{}(t)],$$ (31) where $`\widehat{\alpha }_j,\widehat{\alpha }_j^{}`$ are the bosonic operators satisfying $`[\widehat{\alpha }_i,\widehat{\alpha }_j^{}]=\delta _{ij}`$ in the space of states with non-zero condensate number, and they describe the annihilation or creation of excitations (quasi-particles), or, equivalently, transitions from an excited state $`j`$ into the condensate and vice versa. This implies that the operators $`\widehat{\alpha }_j,\widehat{\alpha }_j^{}`$ do not commute with the condensate operator $`\widehat{a}_0`$. The wave functions $`u_j(𝐫)`$ and $`v_j(𝐫)`$ are the corresponding amplitudes associated with the annihilation of a real particle at position $`𝐫`$, an action which involves both annihilation or creation of excitations on top of the condensate. The time-dependence of the functions $`\psi _0,u_j,v_j`$ is induced only by the change in the global variables $`V_t(𝐫,t),N_t(t),E_{trap}(t)`$ and they are assumed to be time-independent under the steady-state conditions. The condensate wave function $`\psi _0(𝐫)`$ in Eq. (29) is defined as the solution of the generalised steady-state Gross-Pitaevskii equation $$\{_t\mu +U_0[\overline{N}_0|\psi _0(𝐫)|^2+2\overline{n}(𝐫)]\}\psi _0(𝐫)=0$$ (32) where $`_t`$ is given in Eq. (16), while the adiabatic mean number $`\overline{N}_0`$ of the condensate atoms and the density $`\overline{n}(𝐫)`$ of the non-condensate atoms are calculated self-consistently by requiring $$\overline{N}_0+d^3𝐫\overline{n}(𝐫)=N_t.$$ (33) The functions $`u_j(𝐫),v_j(𝐫)`$ satisfy the steady-state equations $`\left(\begin{array}{cc}_t\mu +2U_0[\overline{N}_0|\psi _0(𝐫)|^2+\overline{n}(𝐫)]& U_0\overline{N}_0(\psi _0(𝐫))^2\\ U_0\overline{N}_0(\psi _0^{}(𝐫))^2& _t+\mu 2U_0[\overline{N}_0|\psi _0(𝐫)|^2+\overline{n}(𝐫)]\end{array}\right)\left(\begin{array}{c}u_j(𝐫)\\ v_j(𝐫)\end{array}\right)=E_j\left(\begin{array}{c}u_j(𝐫)\\ v_j(𝐫)\end{array}\right)`$ (40) $`+U_0\overline{N}_0{\displaystyle d^3𝐫|\psi _0(𝐫)|^2[\psi _0^{}(𝐫)u_j(𝐫)+\psi _0(𝐫)v_j(𝐫)]\left(\begin{array}{c}\psi _0(𝐫)\\ \psi _0^{}(𝐫)\end{array}\right)},`$ (43) where $`E_j=\mathrm{}\omega _j`$ are the $`j`$th quasi-particle excitation energy with respect to the condensate ground state energy, and the second term on the right hand side ensures the orthogonality of the non-condensate functions with the condensate. We note that the vectors $`\left(\begin{array}{c}v_j^{}\\ u_j^{}\end{array}\right)`$ satisfy an equation similar to Eq. (43) with $`E_jE_j`$; Eqs. (32) and (43) must be solved self-consistently for any given global conditions. The mean number of atoms in an excited state in equilibrium is assumed to be given by the Bose-Einstein distribution $$n_j^{eq}=\frac{1}{\mathrm{exp}[\mathrm{}\omega _j/T]1}.$$ (44) In this section we assume a weak output process so that the total number of atoms in the trap does not change significantly during the application of the output coupling. Under these conditions the time-dependence of the operators $`\widehat{a}_0`$ and $`\widehat{\alpha }_j`$ is assumed to be simply $`\widehat{a}_0(t)`$ $``$ $`\widehat{a}_0(0)`$ (45) $`\widehat{\alpha }_j(t)`$ $``$ $`\widehat{\alpha }_j(0)e^{i\omega _jt}.`$ (46) In our numerical demonstration throughout this paper we take a one dimensional Bose gas of $`N_t=2000`$ atoms in a harmonic trap with frequency $`\omega `$. The critical temperature for condensation in this case is $`T_c300\mathrm{}\omega /k`$. We have used a self-consistent HFB-Popov method to find the wave functions and energies of the condensate and the excitations. Throughout this paper we take $`U_0=10\mathrm{}\omega \sqrt{2\mathrm{}/m\omega }`$. We present calculations for two temperatures: for $`T=10\mathrm{}\omega /k`$ we obtain the chemical potential $`\mu 2.5\mathrm{}\omega `$ and the non-condensate fraction $`2\%`$. At $`T=150\mathrm{}\omega /k`$ we obtain $`\mu 2.3\mathrm{}\omega /k`$ and the non-condensate fraction is $`44\%`$. We take the interaction strength between the trapped and untrapped atoms to be $`U_1=U_0`$. Length will be presented in units of $`2\sqrt{\mathrm{}/m\omega }`$ (“harmonic-oscillator units”). ### B Basic properties of the output In order to obtain the properties of the output we expand the output field operator $`\widehat{\psi }_f`$ in terms of the free modes $`\phi _𝐤`$ in the quasi-steady-state\[Eq. (21)\]. We assume $`\lambda (𝐫,t)=\lambda (𝐫)e^{i\mathrm{\Delta }_{em}t}`$. Using the form (29) and (31) of $`\widehat{\psi }_t`$ and the assumptions (45) and (46), we obtain from Eq. (24) the following equation for the annihilation operators of the free output modes $`\widehat{b}_𝐤(t)`$ $``$ $`e^{i\omega _𝐤t}\{\widehat{b}_𝐤(0)i\{\lambda _{𝐤0}D_{𝐤0}(t)\widehat{a}_0(0)`$ (48) $`+\widehat{a}_0{\displaystyle \frac{1}{\sqrt{\widehat{N}_0}}}{\displaystyle \underset{j}{}}[\lambda _{𝐤j+}D_{𝐤j+}(t)\widehat{\alpha }_j(0)+\lambda _{𝐤j}D_{𝐤j}(t)\widehat{\alpha }_j^{}(0)]\}\},`$ where $`\lambda _{𝐤0}`$ $`=`$ $`{\displaystyle d^3𝐫\phi _𝐤^{}(𝐫)\lambda (𝐫)\psi _0(𝐫)}`$ (49) $`\lambda _{𝐤j+}`$ $`=`$ $`{\displaystyle d^3𝐫\phi _𝐤^{}(𝐫)\lambda (𝐫)u_j(𝐫)}`$ (50) $`\lambda _{𝐤j}`$ $`=`$ $`{\displaystyle d^3𝐫\phi _𝐤^{}(𝐫)\lambda (𝐫)v_j^{}(𝐫)}`$ (51) are the matrix elements of $`\lambda (𝐫)`$ between the wave functions of the collective excitations and the output states. The time- and energy- dependence is determined by the functions $$D_{𝐤\eta }(t)=i\frac{e^{i(\omega _{out}^\eta \omega _𝐤)t}1}{\omega _{out}^\eta \omega _𝐤},$$ (52) where $$\mathrm{}\omega _{out}^\eta =\mu +\mathrm{\Delta }_{em}+E_\eta ,$$ (53) with $`\eta =0,j+,j`$, and $`E_{j+}=E_j`$ and $`E_j=E_j`$. With the above definitions, the field operator $`\widehat{\psi }_f(𝐫,t)`$ of the free atoms can be written as $`\widehat{\psi }_f(𝐫,t)`$ $``$ $`\widehat{\psi }_f^{(0)}(𝐫,t)i\mathrm{\Psi }_f^0(𝐫,t)\widehat{a}_0(0)`$ (55) $`\widehat{a}_0(0){\displaystyle \frac{i}{\sqrt{\widehat{N}_0}}}{\displaystyle \underset{j}{}}\left[\mathrm{\Psi }_f^{j+}(𝐫,t)\widehat{\alpha }_j(0)+\mathrm{\Psi }_f^j(𝐫,t)\widehat{\alpha }_j^{}(0)\right],`$ where $$\mathrm{\Psi }_f^\eta (𝐫,t)=\underset{𝐤}{}\phi _𝐤(𝐫)D_{𝐤\eta }(t)\lambda _{𝐤\eta }e^{i\omega _𝐤t}$$ (56) for each $`\eta =0,j+,j`$. This result enables us to calculate various properties of the output from the trap, assuming Bose-Einstein statistics for the initial quasiparticle populations inside the trap. For the initial state we assume that all the cross correlation functions between different operators $`\widehat{a}_0,\widehat{\alpha }_j,\widehat{\alpha }_j^{}`$ vanish, and the only non-zero contributions are the populations $`\widehat{a}_0^{}\widehat{a}_0=N_0n_0^t`$ (57) $`\widehat{\alpha }_j^{}\widehat{\alpha }_j=n_jn_{j+}^t`$ (58) $`\widehat{\alpha }_j\widehat{\alpha }_j^{}=n_j+1n_j^t.`$ (59) Any measurable quantities related to the output atoms may be expressed in terms of the correlation functions of the field operator $`\widehat{\psi }_f(𝐫,t)`$ at different times and space points. In particular, the density of output atoms at a given point $`𝐫`$ and time $`t`$ is given by the equal-time, equal-position correlation function $$n_{out}(𝐫,t)=\widehat{\psi }_f^{}(𝐫,t)\widehat{\psi }_f(𝐫,t).$$ (60) By using Eq. (55) and the assumptions (57)-(59) we observe that $`n_{out}(𝐫,t)`$ can be written as a sum over discrete contributions from the levels in the trap $$n_{out}(𝐫,t)=N_0|\mathrm{\Psi }_f^0(𝐫,t)|^2+\underset{j}{}\left[n_j|\mathrm{\Psi }_f^{j+}(𝐫,t)|^2+(n_j+1)|\mathrm{\Psi }_f^j(𝐫,t)|^2\right],$$ (61) where the first term is the condensate output, the second term is the contribution from the stimulated quantum evaporation of the thermal excitations, and the third term is the contribution from the pair-breaking process, as discussed in Ref. . Following similar steps, the number of output atoms in mode $`𝐤`$ of the free atomic level, $`n_𝐤\widehat{b}_𝐤^{}\widehat{b}_𝐤`$, is given by a sum of discrete contributions from the different levels of the trapped gas $$n_𝐤(t)=n_𝐤^0(t)+\underset{j}{}\left[n_𝐤^{j+}(t)+n_𝐤^j(t)\right],$$ (62) where each term $`\eta =0,j+,j`$ in Eq. (62) has the form $$n_𝐤^\eta (t)=|\lambda _{𝐤\eta }|^2|D_{𝐤\eta }(t)|^2n_\eta ^t.$$ (63) The time- and energy- dependence of each of the $`n_𝐤^\eta `$ is given by the function $$|D_{𝐤\eta }(t)|^2=\frac{1}{2}\left[\frac{\mathrm{sin}[(\omega _𝐤\omega _{out}^\eta )t/2]}{(\omega _𝐤\omega _{out}^\eta )/2}\right]^2.$$ (64) This function has a spectral width in $`\omega _𝐤`$ which decreases with time as $`\pi /t`$. This spectral width represents the energy uncertainty dictated by the finite duration of the output coupling process. The time evolution of the output atoms is therefore governed by the spectral dependence of the matrix elements $`\lambda _{𝐤\eta }`$. In order to analyse the evolution of the output rate and the output atom density, we define two frequency scales with regard to the matrix elements $`\lambda _{𝐤\eta }`$ for each $`\eta =0,j+,j`$: (1) $`\mathrm{\Delta }\omega _\eta `$, the frequency bandwidth within which the matrix elements $`\lambda _{𝐤\eta }`$, defined in Eqs. (49)-(51), are significant and (2) $`\delta \omega _\eta `$, the “width” of $`\lambda _{𝐤\eta }`$ in the $`\omega _𝐤\mathrm{}k^2/2m`$ space in the vicinity of $`\omega _𝐤=\omega _{out}^\eta `$. The weak coupling assumption is justified if the strength of the coupling, which may be represented by the parameter $`\mathrm{\Lambda }=\sqrt{d^3𝐫|\lambda (𝐫)|^2}`$ is much smaller than $`\mathrm{\Delta }\omega _\eta `$ of each trap state, namely, $$\mathrm{\Lambda }\mathrm{\Delta }\omega _\eta .$$ (65) If this condition is not satisfied, then we expect Rabi oscillations between the trapped atomic level $`|t`$ and the output level $`|f`$. In the case of weak coupling, we may identify three temporal regimes: 1. Very short times, $`t\mathrm{\Delta }\omega _\eta ^1`$. Then the function $`D_{𝐤\eta }(t)`$ in Eq. (52) becomes independent of $`𝐤`$, and if $`\omega _{out}^\eta `$ lies within the bandwidth $`\mathrm{\Delta }\omega _\eta `$, $`D_{𝐤\eta }(t)t`$. In this case, the completeness of the set of functions $`\phi _𝐤(𝐫)`$ implies $$\mathrm{\Psi }_f^\eta (𝐫,t)_{t0}\lambda (𝐫)\psi _t^\eta (𝐫)t,$$ (66) where $`\psi _t^0=\psi _0,\psi _t^{j+}=u_j`$, and $`\psi _t^j=v_j^{}`$. The initial shape of the output wave functions before it had time to propagate is therefore the overlap of the electromagnetic field amplitude and the corresponding trapped wave function. The density $`n_{out}(𝐫)`$ in this case is then similar in shape to the density of atoms in the trap and the total number of output atoms increases quadratically in time. This result may also be used to calculate the output beam immediately after the application of a strong coupling pulse, before the output beam starts to propagate or Rabi oscillations occur. 2. Intermediate times, $`\mathrm{\Delta }\omega _\eta ^1<t<\delta \omega _\eta ^1`$. In this case the rate of output from each trap state $`\eta `$ may show oscillations, which follow from interference between output from different momentum states. 3. Long times, $`t\delta \omega _\eta ^1`$. The output from the internal state $`\eta `$ is then mainly generated in a narrow range of energies around $`\mathrm{}\omega _{out}^\eta `$ and the rate of output $`dn_𝐤/dt`$ into these specific modes settles on a constant value, which is determined by the absolute value of the matrix element $`\lambda _{𝐤\eta }`$ at $`\omega _𝐤=\omega _{out}^\eta `$. It is then given by the Fermi golden rule $$\frac{dn_𝐤}{dt}=2\pi \underset{\eta }{}n_\eta ^t|\lambda _{𝐤\eta }|^2\delta (\omega _𝐤\omega _{out}^\eta ),$$ (67) and the output rate obtained when one scans the frequency detuning $`\mathrm{\Delta }_{em}`$ of the coupling fields measures the magnitude of the matrix elements $`|\lambda _{𝐤\eta }|^2`$ as a function of $`\omega _𝐤`$. The asymptotic behaviour of $`|D_{𝐤\eta }(t)|^2`$ in this limit is $$|D_{𝐤\eta }(t)|^22\pi \delta (\omega _𝐤\omega _{out}^\eta )t+2\frac{P}{(\omega _𝐤\omega _{out}^\eta )^2},$$ (68) where the principal part in the second term is defined as $`P/(\omega \omega ^\eta )^2/\omega ^\eta [P/(\omega \omega ^\eta )]`$. The first term represents a linearly increasing mono-energetic contribution of the level $`\eta `$ to the output with a constant rate given in Eq. (67). The second term involving the principal part represents a time-independent non-resonant part that has two physical implications. First, it represents those frequency components that are different from the central resonance frequency which arise due to the sudden switching on of the output coupling field at $`t=0`$. Second, this term results in the fields $`\mathrm{\Psi }_f^\eta (t)`$ containing a non-propagating (bound) part $$(\mathrm{\Psi }_f^\eta (𝐫,t))_{bound}=e^{i\omega _{out}^\eta t}P\underset{𝐤}{}\frac{\phi _𝐤(𝐫)\lambda _{𝐤\eta }}{\omega _𝐤\omega _{out}^\eta },$$ (69) which stays mainly near the trap. This term appears as a part of the dressed ground-state of the coupled system, which is a mixture of the trapped and the untrapped atomic levels. It therefore represents a virtual transition to the output level while the atoms remain bound to the trap. It is detectable if the atomic detecting system is sensitive enough to identify small number of atoms in a different Zeeman level near the main atomic cloud, which would contain atoms in the Zeeman level $`|t`$. Although this last contribution is in general small compared to the resonant contributions, it may be significant when considering the condensate output ($`\eta =0`$), which is multiplied by the large number $`N_0`$, and therefore may be dominant near the trap relative to the contributions of the stimulated quantum evaporation (for $`\mathrm{\Delta }_{em}<0`$) or the pair-breaking (for $`\mathrm{\Delta }_{em}>0`$). The condensate contribution can be estimated by $`(n_f^0)_{bound}2N_0(\lambda (0)/\omega _{out}^0)^2`$, where $`\lambda (0)`$ is the Rabi frequency associated with the coupling field at the centre of the trap. The spectral widths $`\mathrm{\Delta }\omega _\eta `$ and $`\delta \omega _\eta `$ defined above may drastically vary with the structure of the Hilbert space of the output modes, which is determined by the potential $`V_f(𝐫)`$, and also with the spatial shape of the coupling $`\lambda (𝐫,t)`$. It is worth mentioning the following limiting cases: 1. In the absence of gravity the wave functions $`\phi _𝐤`$ are roughly given by the plane waves $`e^{i𝐤𝐫}`$. Then the matrix elements $`\lambda _{𝐤0}`$ that couple the condensate to the free modes is roughly the Fourier transform of the condensate wave function $`\psi _0(𝐫)`$. If no momentum kick is provided, then their width in momentum space is given in terms of the spatial width $`r_0`$ of the condensate by $`\delta k_01/r_0`$. The corresponding spectral width, $`\mathrm{\Delta }\omega _0`$, is then $$\mathrm{\Delta }\omega _0\delta \omega _0\mathrm{}/2mr_0^2<\omega ,$$ (70) which implies that the time it will take to achieve a constant rate of output from the condensate is larger than the period of the trap. If the condensate is broadened by a strong collisional repulsion then this time may be much greater than this period, which is typically of the order of $`10`$ ms. 2. In the case where a momentum kick $`𝐤_{em}`$ is provided, the spectral width for the condensate output becomes $$\delta \omega _0\mathrm{}|𝐤_{em}|\delta k_0/2m\mathrm{}|𝐤_{em}|/2mr_0.$$ (71) This makes the time for achieving steady output shorter by a factor $`(|𝐤_{em}|r_0)^1`$ compared to the previous case. If $`|𝐤_{em}|`$ corresponds to an optical wavelength then this factor may be of the order of 10. 3. In the presence of gravity the spectral width of $`\lambda _{𝐤\eta }`$ is determined mainly by the gradient of the gravitational potential over the spatial extent of the corresponding wave function $`\psi _t^\eta `$. A typical value of this gradient for the condensate wave function in the experiment of Ref. is about $`\delta \omega _02\pi \times 10`$ kHz. In this case the time needed to achieve a steady output is much shorter, in the order of $`0.1`$ ms. The rate of transfer of atoms into the output level as a function of $`\mathrm{\Delta }_{em}`$ in our one-dimensional example is plotted in Figure 2. This rate is a sum of contributions from the condensate and the excited states in the trap $$\frac{dN_{out}}{dt}=\frac{dn_f^0}{dt}+\underset{j}{}\left[\frac{dn_f^{j+}}{dt}+\frac{dn_f^j}{dt}\right],$$ (72) where $$n_f^\eta =d^3𝐫n_{out}^\eta (𝐫)=\underset{𝐤}{}n_𝐤^\eta .$$ (73) The rate of output from the condensate component, $`dn_f^0/dt`$ (solid line), that from stimulated quantum evaporation, $`_jdn_f^{j+}/dt`$ (dashed line), and finally from pair-breaking, $`_jdn_f^j/dt`$ (dash-dotted line) are shown for temperatures $`T=10\mathrm{}\omega /k`$ ($`0.03T_c`$, bold line) and $`T=150\mathrm{}\omega /k`$ ($`0.5T_c`$, thin line), in the case where no momentum is transferred from the EM field ($`𝐤_{em}=0`$). The threshold below which the condensate output is not produced is at $`\mathrm{\Delta }_{em}=\mu `$, which is slightly different for the two temperatures. To prevent unphysical effects that follow from the divergence of the density of states at small momenta in one-dimensional systems, we have assumed that the density of momentum states per energy is constant, $`\rho (\omega _𝐤)=1`$. As to be anticipated, the composition of the output beam changes as a function of $`\mathrm{\Delta }_{em}`$: for negative values of $`\mathrm{\Delta }_{em}`$ the dominant contribution is from the stimulated quantum evaporation of initially excited levels in the trap; for positive $`\mathrm{\Delta }_{em}`$ it is the contribution of pair-breaking that dominates the output. The output contains overwhelmingly condensate components at central values of $`\mathrm{\Delta }_{em}`$. Comparison of the results for $`T=10\mathrm{}\omega /k`$ and $`T=150\mathrm{}\omega /k`$ shows that output rate from the pair-breaking is pronounced mainly at low temperatures. Figure 3 is a one-dimensional demonstration of the output density for coupling frequencies in (a) the stimulated quantum evaporation regime ($`\mathrm{\Delta }_{em}=5\omega `$), (b) the coherent output regime ($`\mathrm{\Delta }_{em}=0`$), and (c) the pair-breaking regime ($`\mathrm{\Delta }_{em}=8\omega `$). At very short times the output density from each level has the same shape as the density of the given level in the trap, as follows from Eq. (66). After a short time, the output atoms emerge mainly in two momentum states $`\phi _𝐤`$ corresponding to the right- and left-propagating waves with energy $`\mathrm{}\omega _{out}^\eta `$. Since the magnitude of the matrix elements $`\lambda _{𝐤\eta }`$ for these two modes are equal, the output beam corresponding to a given component $`\eta `$ forms a standing wave and consequently the density $`n_{out}^\eta (x)`$ becomes oscillatory. This aspect is demonstrated below, when we discuss the coherence of the output. These standing waves are not expected in cases where the inversion symmetry is broken, such as the case with $`𝐤_{em}0`$. In cases (a) and (c), where $`\mathrm{\Delta }_{em}`$ is very positive or negative, one can see that the output density from the condensate has a steady component that remains near the trap. This part corresponds to the appearance of the bound states discussed above in conjunction with Eq. (69). ### C Coherence of the output The concept of the $`n`$-th order coherence in a quantum system was originally developed in the optical context to quantify the correlations in the field. The first order coherence measures the fringe contrast in a typical Young’s double slit experiment, while the second order coherence gives indications of counting statistics in, say, Hanbury-Brown-Twiss experiments. A theory of the coherence of matter-waves was presented only recently and the case of a trapped Bose gas has also been discussed. It follows that, for matter waves, the theory of coherence which is applicable to real experiments is much more complicated than that for the optical case. However, any measures of coherence must involve correlation functions between the matter-wave field operators. For simplicity, we use here definitions of matter wave coherence functions that are equivalent to the optical definitions by replacing the usual electromagnetic field operators by the matter-wave field operators. #### 1 First-order coherence The first order coherence function $`g_f^{(1)}(𝐫,𝐫^{},t,t^{})`$ for the output atoms is defined as $$g_f^{(1)}(𝐫,𝐫^{},t,t^{})=\frac{\widehat{\psi }_f^{}(𝐫,t)\widehat{\psi }_f(𝐫^{},t^{})}{\sqrt{\widehat{\psi }_f^{}(𝐫,t)\widehat{\psi }_f(𝐫,t)\widehat{\psi }_f^{}(𝐫^{},t^{})\widehat{\psi }_f(𝐫^{},t^{})}},$$ (74) where $`g^{(1)}=1`$ implies full coherence and $`g^{(1)}=0`$ implies total incoherence. The first-order coherence for a random or thermal mixture of many modes typically takes the maximal value for $`𝐫=𝐫^{}`$ (i.e. $`g^{(1)}(𝐫,𝐫)=1`$) and falls down to zero for large $`|𝐫𝐫^{}|`$ or $`|tt^{}|`$. Highly monochromatic beams, however, are characterised by the fact that $`g^{(1)}=1`$ even for large $`|𝐫𝐫^{}|`$ or $`|tt^{}|`$, implying high fringe visibility even if widely separated parts of the beam interfere. An atomic beam weakly coupled out from a finite temperature Bose-gas is, in general, a mixture of quasi-monochromatic beams originating from the condensate and the internal excitations in the trap. The nature of this mixture depends on the frequency, shape and momentum transfer from the electromagnetic field, and correspondingly the coherence properties are significantly affected. Following the quasi-steady-state solution for $`\widehat{\psi }_f(𝐫)`$ in Eq. (55) we find for our first order coherence, $$g_f^{(1)}(𝐫,𝐫^{},t,t^{})=\frac{1}{\sqrt{n_{out}(𝐫)n_{out}(𝐫^{})}}\underset{\eta =0,j+,j}{}N_\eta [\mathrm{\Psi }_f^\eta (𝐫,t)]^{}\mathrm{\Psi }_f^\eta (𝐫^{},t^{}).$$ (75) The coherence is maximal if only one of the terms from the sum over $`\eta `$ is dominant. Fig. 4 shows the first order coherence function $`g_f^{(1)}(x_1,x_2,t)`$ of the output atoms in our one-dimensional demonstration as a function of $`x_2`$ for fixed $`x_1=0`$ and time $`t=100/\omega `$. When $`\mathrm{\Delta }_{em}=0`$ and the temperature is low ($`T=10\mathrm{}\omega /k`$, Fig. 4a) the coherence function is unity except for points where the condensate density vanishes (see Fig. 3b). At $`T=150\mathrm{}\omega /k`$ (Fig. 4b) the thermal component is larger and it is more dominant near the points where the density of the condensate component is low. These features are unique to configurations where the output has a form of standing matter-waves. When $`\mathrm{\Delta }_{em}=5\omega `$ (Fig. 4c) the thermal components are dominant (see Fig. 3a) and the coherence drops much lower than unity. When $`\mathrm{\Delta }_{em}=8\omega `$ (fig. 4d) only few thermal output components exist from the pair-breaking process, and consequently one obtains a comparatively high coherence function. #### 2 Second-order coherence Of particular interest are the intensity correlations of the fields which are important, for example, in experiments involving non-resonant light scattering from an atomic gas. These intensity correlations are expressed in terms of the second order correlation function, $`g_f^{(2)}(𝐫_1,𝐫_2,t_1,t_2)`$, which is defined as $$g_f^{(2)}(𝐫_1,t_1;𝐫_2,t_2)=\frac{\widehat{\psi }_f^{}(𝐫_1,t_1)\widehat{\psi }_f^{}(𝐫_2,t_2)\widehat{\psi }_f(𝐫_1,t_1)\widehat{\psi }_f(𝐫_2,t_2)}{\widehat{\psi }_f^{}(𝐫_1,t_1)\widehat{\psi }_f(𝐫_1,t_1)\widehat{\psi }_f^{}(𝐫_2,t_2)\widehat{\psi }_f(𝐫_2,t_2)}.$$ (76) The function measures the joint probability of detecting two atoms at two space-time points. If the detection probability of an atom is independent of the detection probability of another atom then $`g^{(2)}=1`$ and the probability distribution is Poissonian. This is the case for a coherent state of the matter field. However, for a thermal state the correlation function at the same space-time point is $`g_f^{(2)}(𝐫_1=𝐫_2,t_1=t_2)2`$. This implies that the atoms are “bunched,” i.e. there is a larger probability to detect two atoms together. The second-order correlation function at equal position and time points $`𝐫_1=𝐫_2`$ and $`t_1=t_2`$ was previously calculated for a trapped Bose gas. Here we shall follow the same treatment for calculating the second-order coherence of the output beam. We decompose the field operator $`\widehat{\psi }_f(𝐫,t)`$ into a part proportional to the condensate and a part proportional to the excited states in the trap, and apply Wick’s theorem to the expectation value of the product of four non-condensate operators: $$\widehat{\psi }_{nc}^{}(𝐫)\widehat{\psi }_{nc}^{}(𝐫)\widehat{\psi }_{nc}(𝐫)\widehat{\psi }_{nc}(𝐫)=2\stackrel{~}{n}^2(𝐫)+\stackrel{~}{m}^{}(𝐫)\stackrel{~}{m}(𝐫),$$ (77) where we have defined $`\stackrel{~}{n}(𝐫)=\widehat{\psi }_{nc}^{}(𝐫)\widehat{\psi }_{nc}(𝐫)`$ and $`\stackrel{~}{m}(𝐫)=\widehat{\psi }_{nc}(𝐫)\widehat{\psi }_{nc}(𝐫)`$. One then obtains for the second order coherence of the output atoms: $$g_f^{(2)}(𝐫,t)=1+\frac{1}{n_{out}^2(𝐫)}\left\{2\mathrm{R}\mathrm{e}\left[n_{out}^0(𝐫)\stackrel{~}{n}_{out}(𝐫)+[(\mathrm{\Psi }_{out}^0(𝐫))^{}]^2\stackrel{~}{m}_{out}(𝐫)\right]+\stackrel{~}{n}_{out}^2(𝐫)+|\stackrel{~}{m}_{out}(𝐫)|^2\right\},$$ (78) where $`\stackrel{~}{n}_{out}(𝐫)=_j[n_{out}^{j+}(𝐫)+n_{out}^j(𝐫)]`$ and $`\stackrel{~}{m}_{out}(𝐫)=_j\mathrm{\Psi }_f^{j+}\mathrm{\Psi }_f^j(2n_j+1)`$. We note that although in Ref. the terms proportional to $`\stackrel{~}{m}`$ had negligible contribution, here they may play an important role even at zero temperature in situations where the tuning of the coupling EM-field frequency yields an output beam that emerges mainly from the non-condensate parts of the trapped gas. The equal-time, single-point intensity correlation of the output atoms after a time $`t=100/\omega `$ in few typical cases is shown in Fig. 5. If the output condensate is dominant (Fig. 5a) the function $`g^{(2)}(x)`$ is equal to unity except at discrete points where the output condensate wave function vanishes. At a higher temperature (Fig. 5b) the thermal output components tend to raise $`g^{(2)}(x)`$ near the points where the coherent part is small. In the case where the thermal component is dominant ($`\mathrm{\Delta }_{em}=5\omega `$, Fig. 5c) $`g^{(2)}(x)`$ assumes the value of 2. In the case where the pair-breaking is dominant (Fig. 5d) the intensity correlations tend to assume values greater than 2. This can be interpreted as an atom bunching effect caused by the combination of the process of pair-breaking with the stimulated quantum evaporation of the thermal states. ## IV Evolution of the trapped gas The last section was devoted to a discussion of the properties of the output in the quasi-steady-state approximation, where the Bose gas in the trap is assumed to remain unchanged by the output coupling process. In this section we describe the internal dynamics of the trapped atomic Bose gas during the output coupling process. During this process, the trapped atomic population of the condensate state and each of the excited states change in a different way and the system is driven out of equilibrium. In the typical case where the duration of the coupling process is short compared to the duration of relaxation processes at very low temperatures, the dynamics is represented by approximate solutions of Eq. (26). In the case of weak coupling, the solutions are best represented in terms of the adiabatic basis of the system, which are the steady-state HFB-Popov solutions for a given total number of particles and given total energy of the system. It can serve as a good basis as long as the changes in the conditions in the trap are slow enough compared with the trap frequency. We begin by first introducing a two-component vector formalism that is convenient for dealing with the many modes of excitations. We then obtain linear equations of motion for the creation and annihilation operators of the condensate and excitations in the adiabatic basis. In the adiabatic conditions these equations may be simplified and solved analytically. A perturbation solution is then presented, which is suitable for describing the short time evolution. Finally, we find that the number-conserving formalism fails to describe the evolution of the condensate number in the pair-breaking regime. This problem is discussed and cured in the end of this section. ### A Vector formalism for the trapped atomic gas The dynamics of the excited states in the trap is usually described by a set of two coupled equations of the form Eq. (43), which was discussed in Sec. III, or its time-dependent version. This form, as well as the fact that the expansion for the field operator in Eq. (29) involves the quasiparticle creation and annihilation operators $`\widehat{\alpha }_j,\widehat{\alpha }_j^{}`$, motivates the introduction of the two-component vector formalism as follows. First, we define the normalised condensate operators $$\widehat{c}_0=\widehat{a}_0\frac{1}{\sqrt{\widehat{N}_0}};\widehat{c}_0^{}=\frac{1}{\sqrt{\widehat{N}_0}}\widehat{a}_0^{},$$ (79) which are well-defined in the space spanned by states with non-zero condensate number; within this space, they satisfy $`\widehat{c}_0\widehat{c}_0^{}=\widehat{c}_0^{}\widehat{c}_0=1`$. We then define the two-component column vector operator $$\widehat{\xi }_t(𝐫,t)=\left(\begin{array}{c}\widehat{c}_0^{}\widehat{\psi }_t(𝐫,t)\\ \widehat{\psi }_t^{}(𝐫,t)\widehat{c}_0\end{array}\right)$$ (80) which describes transitions from the condensate state to itself and to and from the excited states. The expansion of $`\widehat{\psi }_t(𝐫,t)`$ in Eq. (29) and (31) is equivalent to the expansion of $`\widehat{\xi }_t(𝐫,t)`$ in terms of the two-component wave function vectors as $$\widehat{\xi }_t(𝐫,t)=\underset{\eta =\mathrm{}}{\overset{\mathrm{}}{}}\xi _\eta (𝐫,t)\widehat{\alpha }_\eta (t)e^{i_0^t𝑑t^{}E_\eta (t^{})},$$ (81) where the index $`\eta `$ takes integer values from $`\mathrm{}`$ to $`\mathrm{}`$. The index $`\eta =0`$ corresponds to the condensate; negative indices stand for solutions of Eq. (43) with negative energies $`E_j=E_j`$, and operators such that $`\widehat{\alpha }_j=\widehat{\alpha }_j^{}`$. In addition, we note that the condensate operator $$\widehat{\alpha }_0\widehat{c}_0^{}\widehat{a}_0=\sqrt{\widehat{N}_0}$$ (82) is Hermitian. The two-component vectors $`\xi _\eta (𝐫,t)`$ are defined as $$\xi _0\left(\begin{array}{c}\psi _0\\ \psi _0^{}\end{array}\right),\xi _j\left(\begin{array}{c}u_j\\ v_j\end{array}\right),\xi _j\left(\begin{array}{c}v_j^{}\\ u_j^{}\end{array}\right).$$ (83) The time dependence of the vectors $`\xi _\eta (𝐫,t)`$ and the energies $`E_\eta (t)`$ in Eq. (81) is governed by the time-dependence of the global variables of the system, while the time dependence of the coefficients $`\widehat{\alpha }_\eta `$, $`\eta (\mathrm{},\mathrm{})`$, represents changes in the populations of the condensate and the excited states. The usual orthogonality and normalisation conditions for the eigenfunction $`u_j`$ and $`v_j`$ are written in the vectorised notation as $$d^3𝐫\xi _0^{}(𝐫)\xi _\eta (𝐫)=d^3𝐫\xi _0^{}(𝐫)\sigma _3\xi _\eta (𝐫)=0,$$ (84) $$d^3𝐫\xi _\eta ^{}(𝐫)\sigma _3\xi _\nu (𝐫)=\mathrm{sign}\{E_\eta \}\delta _{\eta \nu },$$ (85) for any $`\eta ,\nu 0`$. Here $$\xi _j^{}(u_j^{}v_j^{});;\xi _j^{}(v_ju_j)$$ (86) are the two component row vectors and $`\sigma _3`$ denotes a $`2\times 2`$ matrix, $$\sigma _3=\left(\begin{array}{cc}1& 0\\ 0& 1\end{array}\right).$$ (87) ### B Equations of motion for the operators We now derive the equations of motion for the operators $`\widehat{\alpha }_\eta `$ corresponding to transitions from the condensate to the adiabatic eigenmodes of the system and vice versa. We first multiply Eq. (26) by $`\widehat{c}_0^{}e^{i\mathrm{\Phi }}`$. The resulting equation, together with its Hermitian conjugate, form a set of equations which can be expressed in the following vector form $`\dot{\widehat{\xi }}_t`$ $`=`$ $`(\dot{\widehat{\xi }}_t)^{(0)}{\displaystyle _0^t}𝑑t^{}{\displaystyle d^3𝐫^{}e^{i\sigma _3[\mathrm{\Phi }(t)\mathrm{\Phi }(t^{})]}\stackrel{~}{G}(𝐫,𝐫^{},t,t^{})\widehat{\xi }_t(𝐫^{},t^{})}`$ (89) $`i\stackrel{~}{\lambda }^{}(𝐫)\sigma _3\widehat{\xi }_f^{(0)}(𝐫)`$ where $`\stackrel{~}{G}(𝐫,𝐫^{},t,t^{})`$ and $`\stackrel{~}{\lambda }(𝐫)`$ are the matrices $$\stackrel{~}{G}(𝐫,𝐫^{},t,t^{})=\left(\begin{array}{cc}\widehat{c}_0^{}(t)G(𝐫,𝐫^{},t,t^{})\widehat{c}_0(t^{})& 0\\ 0& \widehat{c}_0^{}(t^{})G^{}(𝐫,𝐫^{},t,t^{})\widehat{c}_0(t)\end{array}\right),$$ (90) $$\stackrel{~}{\lambda }=\left(\begin{array}{cc}\lambda (𝐫)& 0\\ 0& \lambda ^{}(𝐫)\end{array}\right),$$ (91) and in a similar manner to $`\widehat{\xi }_t(𝐫,t)`$ of Eq. (80), $$\widehat{\xi }_f^{(0)}(𝐫,t)=\left(\begin{array}{c}\widehat{c}_0^{}\widehat{\psi }_f^{(0)}(𝐫,t)\\ (\widehat{\psi }_f^{(0)})^{}(𝐫,t)\widehat{c}_0\end{array}\right).$$ (92) $`\widehat{\xi }_f^{(0)}(𝐫,t)`$ describes the free evolution of the output field operator $`\widehat{\psi }_f(𝐫,t)`$, as given in Eqs. (19), (22). The term $`(\dot{\widehat{\xi }}_t)^{(0)}`$ is the operator describing the free evolution of $`\widehat{\xi }_t`$ inside the trap in the absence of the output coupling but with a given adiabatic change in the global variables. Here we use the same approximations as in Eqs. (45), (46), which is equivalent to $$(\dot{\widehat{\xi }}_t)^{(0)}(𝐫,t)=i\underset{\eta }{}E_\eta \xi _\eta \widehat{\alpha }_\eta e^{i_0^tE_\eta (t^{})𝑑t^{}}.$$ (93) The time derivative of $`\dot{\widehat{\xi }}_t`$ on the left hand side of Eq. (89) may then be written as $$\dot{\widehat{\xi }}_t=\underset{\eta }{}e^{i_0^t𝑑t^{}E_\eta (t^{})}[\dot{\xi }_\eta \widehat{\alpha }_\eta +\xi _\eta \dot{\widehat{\alpha }}_\eta iE_\eta \xi _\eta \widehat{\alpha }_\eta ].$$ (94) The first term corresponds to the time dependence due to the change in the global variables, the second term is due to the change in the populations of the condensate and excited states, while the third term cancels with $`(\dot{\widehat{\xi }}_t)^{(0)}`$ on the right-hand side of Eq. (89). We multiply Eq. (89), in turn, by $`\xi _\eta ^{}\sigma _3`$ for every $`\eta 0`$ and by $`\frac{1}{2}\xi _0`$ for $`\eta =0`$, and integrate over $`𝐫`$. This multiplication should be understood as an inner product between row and column vectors. By applying the orthogonality and normalisation relations in Eq. (84) and Eq. (85) we obtain the required equation of motion $$\dot{\widehat{\alpha }}_\eta =\underset{\nu }{}M_{\eta \nu }(t)\widehat{\alpha }_\nu (t)\underset{\nu }{}_0^t𝑑t^{}G_{\eta \nu }(t,t^{})\widehat{\alpha }_\nu (t^{})id^3𝐫F_\eta (𝐫,t)\widehat{\xi }_f^{(0)}(𝐫,t).$$ (95) Here $$M_{\eta \nu }(t)=e^{i_0^t𝑑t^{}[E_\eta (t^{})E_\nu (t^{})]}d^3𝐫\xi _\eta ^{}(𝐫)\sigma _\eta \dot{\xi }_\nu (𝐫)$$ (96) is a matrix with zero diagonal, which describes mixing between the adiabatic levels that is induced by the change in the global variables. This term in Eq. (95) may be neglected in the adiabatic limit where the change in the global variables is very slow. Its effect in slightly non-adiabatic conditions will be discussed elsewhere. The second and third terms in Eq. (95) describe changes in the trap which are directly induced by the output coupling. Defining $$\mathrm{\Phi }_\eta (t)=_0^t𝑑t^{}[\mu (t^{})+\sigma _3E_\eta (t^{})],$$ (97) and $$\sigma _\eta =\{\begin{array}{cc}\sigma _3& \eta >0\\ \frac{1}{2}& \eta =0\\ \sigma _3& \eta <0\end{array},$$ (98) one may write $$G_{\eta \nu }(t,t^{})=d^3𝐫d^3𝐫^{}\xi _\eta ^{}(𝐫,t)\sigma _\eta \stackrel{~}{G}(𝐫,𝐫^{},t,t^{})e^{i\sigma _3[\mathrm{\Phi }_\eta (t)\mathrm{\Phi }_\nu (t^{})]}\xi _\nu (𝐫^{},t^{}),$$ (99) and $$F_\eta (𝐫)=\xi _\eta ^{}(𝐫)\sigma _\eta \sigma _3e^{i\sigma _3\mathrm{\Phi }_\eta }\stackrel{~}{\lambda }^{}(𝐫)$$ (100) which describes the effect of the zero-field fluctuations. An exact analytical solution of Eq. (95) is, in general, not possible. However, in the following we present two methods of approximate solutions to this equation: an adiabatic approximation, which is suitable for describing the evolution at long enough times, and a perturbative expansion, which is suitable for short times. ### C Solution via adiabatic approximation First we consider the adiabatic and quasi-continuous case where the functions $`\xi _\eta `$ change very slowly with time and the coupling amplitude is given by $`\lambda (𝐫,t)=\lambda (𝐫)e^{i\mathrm{\Delta }_{em}t}`$. In this case we let $`M_{\eta \nu }0`$, Second, Eq. (95) is further simplified by finding an approximate expression for the integral involving $`G_{\eta \nu }(t,t^{})`$. We make a Markovian approximation, which transforms the integro-differential equation (95) into an ordinary differential equation, which can then be solved analytically. Following the definition of $`G(𝐫,𝐫^{},t,t^{})`$ in terms of the free output modes denoted by $`𝐤`$ \[Eq. (23) and (27)\], the functions $`G_{\eta \nu }(t,t^{})`$ may be written as $$G_{\eta \nu }(t,t^{})=\underset{𝐤}{}\overline{\lambda }_{𝐤\eta }^{}(t)\sigma _\eta e^{i[\omega _𝐤\mathrm{\Delta }_{em}]\sigma _3(tt^{})}e^{i[\mathrm{\Phi }_\eta (t)\mathrm{\Phi }_\nu (t^{})]}\overline{\lambda }_{𝐤\nu }(t^{}),$$ (101) where $$\overline{\lambda }_{𝐤\eta }(t)=\left(\begin{array}{c}\lambda _{𝐤\eta }\\ \lambda _{𝐤,\eta }^{}\end{array}\right),$$ (102) with the matrix element $`\lambda _{𝐤\eta }`$ as defined in Eqs. (49)-(51). The time dependence of the matrix elements $`\lambda _{𝐤\eta }`$ is induced only by the change in the global variables, which is assumed to be slow. The sum over $`𝐤`$ in Eq. (101) may then be regarded as a Fourier transform of the products $`\overline{\lambda }_{𝐤\eta }^{}\sigma _\eta \overline{\lambda }_{𝐤\nu }`$ over $`\omega _𝐤`$ at “fixed” point in time $`\tau tt^{}`$. The width $`\mathrm{\Delta }\tau _{\eta \nu }`$ of $`G_{\eta \nu }(\tau )`$ as a function of $`\tau `$ is then roughly given by the inverse of the spectral width $`\mathrm{\Delta }\omega _{\eta \nu }`$ of the product of the matrix elements $`\lambda _{𝐤\eta }`$ and $`\lambda _{𝐤\nu }`$, which is, in turn, given by the smallest of the spectral widths $`\mathrm{\Delta }\omega _\eta `$ and $`\mathrm{\Delta }\omega _\nu `$ of the corresponding matrix elements. In the same conditions that allow the weak coupling approximations done in Eq. (67), i.e., when $`G_{\eta \nu }\mathrm{\Delta }\tau _{\eta \nu }1`$ and $`t\mathrm{\Delta }\tau _{\eta \nu }`$, we may take $`\widehat{\alpha }_\eta (t^{})\widehat{\alpha }_\eta (t)`$ in Eq. (95) and $`\widehat{c}_0^{}(t)\widehat{c}_0(t^{})=1`$, and extend the integration over $`t^{}`$ to $`\mathrm{}`$, namely $$_0^t𝑑t^{}G_{\eta \nu }(t,t^{})\widehat{\alpha }_\nu (t^{})\mathrm{\Gamma }_{\eta \nu }(t)e^{i_0^t𝑑t^{}[E_\eta (t^{})E_\nu (t^{})]}\widehat{\alpha }_\nu (t),$$ (103) where $`\mathrm{\Gamma }_{\eta \nu }(t)`$ $`=`$ $`{\displaystyle _0^{\mathrm{}}}𝑑\tau {\displaystyle \underset{𝐤}{}}\overline{\lambda }_{𝐤\eta }^{}(t)\sigma _\eta \mathrm{exp}\{i\sigma _3[\omega _𝐤\mu \sigma _3E_\nu \mathrm{\Delta }_{em}i\sigma _3ϵ]\tau \}\overline{\lambda }_{𝐤\nu }(t)`$ (104) $`=`$ $`i{\displaystyle \underset{𝐤}{}}\overline{\lambda }_{𝐤\eta }^{}(t)\sigma _\eta \sigma _3{\displaystyle \frac{1}{\omega _𝐤\mu \sigma _3E_\nu \mathrm{\Delta }_{em}i\sigma _3ϵ}}\overline{\lambda }_{𝐤\nu }(t).`$ (105) The complex fraction should be understood as $`{\displaystyle \frac{i}{xiϵ}}=\pm \pi \delta (x)i{\displaystyle \frac{P}{x}},`$ where $`P/x`$ means the principal part of $`1/x`$ when integrating over $`x`$. Further simplification is achieved when we notice that if the terms $`\mathrm{\Gamma }_{\eta \nu }`$ are much smaller than the energy splittings $`E_\eta E_\nu `$ between the excitation levels in the trap, then the cross-terms with $`\eta \nu `$ oscillate as fast as $`e^{i(E_\eta E_\nu )t}`$ and their contribution averages to zero. We then obtain a system of separate uncoupled equations for each operator $`\widehat{\alpha }_\eta `$, which is given, for non-negative $`\eta =j0`$, by $$\dot{\widehat{\alpha }}_j=\mathrm{\Gamma }_{jj}(t)\widehat{\alpha }_j(t)id^3𝐫F_j(𝐫,t)\widehat{\xi }_f^{(0)}(𝐫,t),$$ (106) for which the solution is $$\widehat{\alpha }_j(t)=\mathrm{exp}[_0^t\mathrm{\Gamma }_{jj}(t^{})𝑑t^{}]\widehat{\alpha }_j(0)i_0^t𝑑t^{}d^3𝐫F_j(𝐫,t^{})e^{_t^{}^t\mathrm{\Gamma }_{jj}(t^{\prime \prime })𝑑t^{\prime \prime }}\widehat{\xi }_f^{(0)}(𝐫,t^{}).$$ (107) For $`j=0`$, $$\mathrm{\Gamma }_{00}(t)=\pi \underset{𝐤}{}|\lambda _{𝐤0}|^2\delta (\omega _𝐤\mu \mathrm{\Delta }_{em}),$$ (108) and for $`j0`$, $$\mathrm{\Gamma }_{jj}=\mathrm{\Gamma }_{j+}+\mathrm{\Gamma }_j$$ (109) where $$\mathrm{\Gamma }_{j\pm }=i\underset{𝐤}{}\frac{|\lambda _{𝐤j\pm }|^2}{\omega _𝐤\mu E_j\mathrm{\Delta }_{em}iϵ}.$$ (110) The imaginary part of $`\mathrm{\Gamma }_{jj}`$ represents energy shifts induced by the output coupling, while its real part $`\gamma _j\mathrm{Re}\mathrm{\Gamma }_{jj}`$ is given by $$\gamma _{j\pm }\mathrm{Re}\mathrm{\Gamma }_{j\pm }=\pm \underset{𝐤}{}|\lambda _{𝐤j\pm }|^2\delta (\omega _𝐤\mu E_j\mathrm{\Delta }_{em})$$ (111) representing decay ($`\gamma _{j+}>0`$) or growth ($`\gamma _j<0`$) of the population of excited level $`j`$. We now proceed to calculate the number of condensate and quasi-particle excitations inside the trap under the adiabatic approximation. The evolution of the condensate number is given straight-forwardly by $$N_0(t)=\widehat{\alpha }_0^2=N_0(0)e^{2\gamma _0t}.$$ (112) However, for calculating $`n_j(t)=\widehat{\alpha }_j^{}(t)\widehat{\alpha }_j(t)`$ we must also consider the free term in Eq. (107), whose contribution is proportional to the correlations of the free field operators $`\widehat{\psi }_f^{(0)}(𝐫,t)`$ $$\widehat{\psi }_f^{(0)}(𝐫,t)(\widehat{\psi }_f^{(0)})^{}(𝐫^{},t^{})=\underset{𝐤}{}\phi _𝐤(𝐫)e^{i\omega _𝐤(tt^{})}\phi _𝐤^{}(𝐫^{})=K_f(𝐫,𝐫^{},tt^{}).$$ (113) In the case of very weak coupling, where $`\mathrm{\Gamma }_{jj}`$ may be assumed to be time-independent, the contribution of this last term in Eq. (107) to $`n_j(t)`$ is $`n^{(0)}(t)`$ $`=`$ $`{\displaystyle \underset{𝐤}{}}|\lambda _{𝐤j}|^2{\displaystyle \frac{|e^{\gamma _jt}e^{i(\omega _𝐤\overline{\omega }_{out}^j)t}|^2}{(\omega _𝐤\overline{\omega }_{out}^j)^2+\gamma _j^2}}`$ (114) $``$ $`2{\displaystyle 𝑑\omega \underset{𝐤}{}|\lambda _{𝐤ȷ}|^2\delta (\omega \omega _𝐤+\overline{\omega }_{out}^j)\frac{|e^{\gamma _jt}e^{i\omega t}|^2}{\omega ^2+\gamma _j^2}},`$ (115) where $`\overline{\omega }_{out}^j=\omega _{out}^j+\mathrm{Im}\mathrm{\Gamma }_{jj}`$. The spectral width of the integrand is $`\mathrm{\Delta }\omega \pi /t`$ for $`|\gamma _jt|1`$ and $`\mathrm{\Delta }\omega \gamma _j`$ for $`|\gamma _jt|1`$. Under the conditions that led to Eq. (106), one may take $`\omega 0`$ in the $`\delta `$-function and consequently identify $`_𝐤|\lambda _{𝐤ȷ}|^2\delta (\omega _𝐤\overline{\omega }_{out}^j)=\gamma _j`$. The integration over $`\omega `$ may be then performed to give the final result $$n_j(t)=\mathrm{exp}[2\gamma _jt]n_j(0)2\gamma _j\frac{1e^{2\gamma _jt}}{2\gamma _j}.$$ (116) This equation is the solution of the differential equation $$\frac{dn_j}{dt}=2\gamma _{j+}n_j(t)2\gamma _j[n_j(t)+1].$$ (117) Here, the first term on the right-hand-side is responsible for an exponential decrease in the number of excitations due to stimulated quantum evaporation, while the second term is responsible for an exponential increase in the number of excitations due to the process of pair breaking, which may start even when the excited states are initially unpopulated. This increase in the number of excitations must, quite clearly, lead to the increase in the number of atoms in the excited states, together with an increase in the number of output atoms. There is, however, no process that may balance this growth in the total number of atoms, and this implies that the growth must be compensated by a decrease in the number of condensate atoms. This is, in fact, not evident from the above equations and the problem is discussed at the end of this section. ### D Solution via perturbation theory A full solution of the linear integro-differential equations Eq. (95) may be sought by perturbative iterations, taking the magnitude of the coupling strength $`\lambda `$ as a perturbative small parameter. Here we present the second-order perturbative solutions, which are valid at short times when the population in different excitation levels are not changed significantly from their initial value. In this case we may also assume that the wave functions and energies of the condensate and excitations are not changed significantly from their initial values (i.e. $`M_{\eta \nu }0`$). If we take the zeroth order solution to Eq. (95) to be given by Eqs. (45) and (46), then the second order solution is given by $`\widehat{\alpha }_\eta (t)`$ $`=`$ $`{\displaystyle \underset{\nu }{}}\left\{\delta _{\eta \nu }{\displaystyle _0^t}𝑑t^{}{\displaystyle _0^t^{}}𝑑t^{\prime \prime }G_{\eta \nu }(t^{},t^{\prime \prime })\right\}\widehat{\alpha }_\nu (0)`$ (119) $`i{\displaystyle _0^t}𝑑t^{}{\displaystyle d^3𝐫F_\eta (𝐫,t)\widehat{\xi }_f^{(0)}(t)}.`$ Under the above assumption, we may perform the integration to obtain $`\widehat{\alpha }_\eta (t)`$ $`=`$ $`{\displaystyle \underset{\nu }{}}\left\{\delta _{\eta \nu }{\displaystyle \underset{𝐤}{}}\overline{\lambda }_{𝐤\eta }^{}\sigma _\eta \overline{D}_{𝐤\eta \nu }^{(2)}(t)\overline{\lambda }_{𝐤\nu }\right\}\widehat{\alpha }_\nu (0)`$ (123) $`i{\displaystyle \underset{𝐤}{}}\overline{\lambda }_{𝐤\eta }^{}\sigma _\eta \sigma _3\overline{D}_{𝐤\eta }(t)\left(\begin{array}{c}\widehat{c}_0^{}\widehat{b}_𝐤\\ \widehat{b}_𝐤^{}\widehat{c}_0\end{array}\right).`$ Here $$\overline{D}_{𝐤\eta }=\left(\begin{array}{cc}D_{𝐤\eta }& 0\\ 0& D_{𝐤,\eta }^{}\end{array}\right),$$ (124) where the functions $`D_{𝐤\eta }`$ are defined in Eq. (52) and $`\overline{D}_{𝐤\eta \nu }^{(2)}(t)`$ $`=`$ $`\{\begin{array}{cc}\frac{i}{E_\eta E_\nu }\left[D_{𝐤\eta }(t)e^{i(E_\eta E_\nu )t}D_{𝐤\nu }(t)\right]& \eta \nu \\ \frac{1i\{\sigma _3[\omega _𝐤\mathrm{\Delta }_{em}\mu ]E_\eta \}te^{i\{\sigma _3[\omega _𝐤\mathrm{\Delta }_{em}\mu ]E_\eta \}t}}{\{\sigma _3[\omega _𝐤\mathrm{\Delta }_{em}\mu ]E_\eta \}^2}& \eta =\nu .\end{array}`$ (127) By using the identity $$|D_{𝐤\eta }(t)|^2=2\mathrm{R}\mathrm{e}\{D_{𝐤\eta \eta }^{(2)}(t)\},$$ (128) \[see Eq. (64)\] we obtain the following expression for the number of condensate atoms in the trap $$N_0(t)=\widehat{\alpha }_0^2(t)=N_0(0)\left[1\underset{𝐤}{}|\lambda _{𝐤0}|^2|\overline{D}_{𝐤0}(t)|^2\right],$$ (129) and for the population of the excited levels we obtain $`n_j(t)`$ $`=`$ $`n_j(0)\left\{1{\displaystyle \underset{𝐤}{}}\left[|\lambda _{𝐤j+}|^2|D_{𝐤j+}(t)|^2|\lambda _{𝐤j}|^2|D_{𝐤j}(t)|^2\right]\right\}`$ (131) $`+{\displaystyle \underset{𝐤}{}}|\lambda _{𝐤j}|^2|D_{𝐤j}(t)|^2.`$ Comparison of Eqs. (129), (131) with the equivalent expressions for the number of output atoms in Eqs. (62), (63) shows that exactly one condensate particle is taken out of the trap per each output atom generated by the coherent output process, while one excitation (quasi-particle) is taken from the trap per each output atom generated by the stimulated quantum evaporation, and one excitation (quasi-particle) is created per each atom that leaves the trap through the pair-breaking process. From Eq. (123) it is straightforward to compute the correlations $`\widehat{\alpha }_\eta ^{}(t)\widehat{\alpha }_\nu (t)`$ between the condensate and the excited levels and between the different levels in the trap. However, it may be shown that only diagonal terms $`\eta =\nu `$ are growing in magnitude with time, while the off-diagonal correlations remain small even after a long time and represent the effects of mixing between different levels induced by the coupling interaction. ### E Number of particles and energy The above treatment of the evolution of the system of a trapped Bose gas has used a formalism which conserves the total number of particles in the system. However, Eqs. (112), (129) show that the change in the number of condensate atoms in the system is independent of the changes in the number of quasi-particles in the trap. This leads to an apparent violation of number conservation; this violation is most pronounced in the process of pair-breaking in which output atoms are created together with quasi-particles in the trap. The problem arises because we ignored the off-diagonal part in the Hamiltonian which is also responsible for the changes in the number of condensate atoms, i.e. $$\widehat{}_{\mathrm{off}\mathrm{diag}}=U_0d^3𝐫\left(\psi _0^{}(𝐫)\right)^2\widehat{a}_0^{}\widehat{a}_0^{}\widehat{\psi }_{nc}(𝐫)\widehat{\psi }_{nc}(𝐫)+H.c.$$ (132) This part of the Hamiltonian, which is responsible for the generation of quantum entanglement between the condensate and the excited states, is washed-out in any mean-field treatment such as the HFB-Popov treatment used here. In the mean-field theory the time-evolution of the condensate operator $`\widehat{a}_0`$ in the steady-state is simply given by Eq. (45), and this leads to the apparent violation of number conservation when the number of quasi-particles in the system is changing. A rigorous theory which corrects this fault is beyond the scope of this paper. Such a theory is in principle straightforward, but technically a little complex: we have to incorporate the anomalous average $`\widehat{\psi }_{nc}(𝐫)\widehat{\psi }_{nc}(𝐫)`$ into the calculation of the condensate wave function and show how this anomalous average acquires an imaginary part in the presence of output coupling of excited states. This, from another viewpoint, represents the change in the effective $`T`$-matrix for the interaction potentials in the presence of decay. Here, we will incorporate number-conservation by requiring that the number of condensate atoms $`N_0(t)`$ is to be determined from the conservation of the total number of particles. If the evolution is adiabatic then some time after the switching-on of the coupling interaction the mixing between different quasi-particle levels may be neglected. The total number of atoms in the trap is then given by $$N_t(t)=N_0(t)+\underset{j}{}\left\{n_j(t)d^3𝐫[|u_j(𝐫)|^2+|v_j(𝐫)|^2]+d^3𝐫|v_j(𝐫)|^2\right\}.$$ (133) On the other hand, we must require $$N_t(t)=N_t(0)N_{out}(t).$$ (134) If we compare Eqs. (62), (63) with Eqs. (129), (131) we see that in the process of stimulated quantum evaporation ($`\eta =j+`$) the number of quasi-particles in the trap decreases in the same rate as the number of output atoms increases, while in the pair-breaking process ($`\eta =j`$) the number of quasi-particles in the trap increases in the same rate as the number of output atoms increases. In other words, in the stimulated quantum evaporation process one thermal quasi-particle is transformed into a real output atom, while in the pair-breaking process one quasi-particle is generated per each output atom that leaves the trap. From inspection of Eq. (133), this implies that for each atom that leaves the trap in a stimulated quantum evaporation(SQE) process, the number of particles associated with the quasi-particle $`j`$ in the trap decreases as $$\delta N_j^{SQE}=d^3𝐫[|u_j(𝐫)|^2+|v_j(𝐫)|^2]=12d^3𝐫|v_j(𝐫)|^2.$$ (135) This must be compensated by an increase in the condensate atom number by $$\delta N_0^{SQE}=+2d^3𝐫|v_j(𝐫)|^2.$$ (136) On the other hand, in the pair-breaking (PB) process, the number of particles associated with the quasi-particle $`j`$ in the trap increases by $$\delta N_j^{PB}=1+2d^3𝐫|v_j(𝐫)|^2.$$ (137) This must be compensated by a decrease in the condensate atom number by $$\delta N_0^{PB}=2d^3𝐫|u_j(𝐫)|^2=2\left(1+d^3𝐫|v_j(𝐫)|^2\right).$$ (138) These considerations lead us to corrections to Eq. (112) which contains only the changes in the condensate particles originating from direct output from the condensate component of the Bose gas. The rate equation for the condensate atoms is now $$\frac{dN_0}{dt}=2\gamma _0N_02\underset{j}{}\{d^3𝐫|v_j(𝐫)|^2|\frac{dn_j^{SQE}}{dt}+d^3𝐫|u_j(𝐫)|^2\frac{dn_j^{PB}}{dt}\},$$ (139) where $`dn_j^{SQE}/dt`$ and $`dn_j^{PB}/dt`$ are the first and second terms on the right-hand side of Eq. (117). The solution of Eq. (139) should now replace the previous solution for $`N_0(t)`$ in Eq. (112). The plots of the time evolution of the trapped condensate and non-condensate populations for few temperatures and coupling parameters are given in Fig. 6. These plots are solutions of the differential equations (117) and (139). When $`\mathrm{\Delta }_{em}=0`$ (Fig. 6a,b) the condensate part decreases while the thermal part does not change significantly. When $`\mathrm{\Delta }_{em}=5\omega `$ (Fig. 6c) conservation of energy only permits transitions from upper excited states to the output level, and the population of the condensate and the lower excited states thus remains unchanged. The upper excited states, in this case, are found to depopulate completely as to be expected. When $`\mathrm{\Delta }_{em}=8\omega `$ (Fig. 6d) the thermal population grows significantly due to transitions of unpaired atoms from the condensate into the excited states. However, the energy distribution in the lower excited states is a highly non-equilibrium distribution and dissipation and thermalization effects that have not been taken into account in this paper should play a major role. The short time limit i.e. $`0t<10\omega ^1`$ behaviour is clearly not accurately described in these plots but it can be calculated by using the low-order perturbative expansions of Sec. IV D. Changes in the total energy in the trap may be caused either by the transfer of atoms out of the trap or by the changes in the chemical potential $`\mu `$ and energies $`E_j`$ of the excitations. The second kind of process is beyond the scope of this paper, since we have neglected changes in $`\mu `$ and $`E_j`$ and put $`M_{\eta \nu }=0`$ in Eq. (95). As for the first kind of process, an energy quantum of $`\delta E=\mu `$ leaves the trap for each condensate atom that leaves the trap (consequently the energy in the trap is reduced by $`\mu `$), the energy changes by $`\delta E_j^{SQE}=\mu E_j`$ for each atom that leaves the trap by the stimulated quantum evaporation process, and finally $`\delta E_j^{PB}=\mu +E_j`$ for each atom that leaves the trap through the pair-breaking process. Therefore we have a relatively simple result for the rate of change in energy: $$\frac{dE_t}{dt}=\mu \frac{dN_t}{dt}+\underset{j}{}E_j\frac{dn_j}{dt}.$$ (140) ## V Summary and Discussion In this paper we have set up a general theory of weak output coupling from a trapped Bose-Einstein gas at finite temperatures. The formalism developed here is suitable for the discussion of both Radio-frequency or stimulated Raman output couplers. It has enabled us to gain much information on the basic features that we expect in real experiments: the time-dependence of the output beam, the effects of excitations in the trapped Bose gas and the pairing of particles. Predictions for specific systems can also be based on our theory. For the time-dependence of the output beam, we have shown that the output beam is a mixture of components from different origins in the trap. The output condensate ($`\eta =0`$) is the coherent part of the beam, while each excited level $`j`$ in the trap contributes two partial waves: one originating from the process of stimulated quantum evaporation ($`\eta =j+`$), where a quasi-particle (excitation) in the trap transforms into a real output atom, and the other originating from the pair-breaking process ($`\eta =j`$), where two correlated atoms in the trap transform into a quasi-particle in the trap and a real output atom. We have shown that a steady monochromatic wave from each component is formed after a time comparable to the inverse of the bandwidth of the corresponding matrix element $`\lambda _{𝐤\eta }`$ as a function of $`\omega _𝐤`$. We have also analyzed the oscillatory behaviour of the output rate at short times and showed the existence of non-propagating bound states in the untrapped level that are formed near the trap as a result of the mixing induced by the output coupler. As for the evolution of the Bose gas in the trap during the process of output coupling, we have shown that for the case of weak coupling an adiabatic approximation may be made, which enables calculation of the composition of the Bose gas inside the trap in terms of the adiabatic basis of condensate and excitations. We have shown that exponential decay of the excitations is expected when the stimulated quantum evaporation process is dominant, while an exponential growth of the number of excitations is expected when the pair-breaking process is dominant. We have shown that the number of trapped condensate atoms increases in each event of stimulated quantum evaporation, while it decreases by more than 2 atoms per each event of pair-breaking. However, we stress that a more elaborate number-conserving theory of time-dependent evolution of the Bose gas in an open system than that considered here is needed. The coherence of the output beam was shown to depend on parameters under experimental control such as the detuning of the laser. We note that the coherence of the output atoms also tells us about the coherence properties of atoms inside the trap; the coherence of the trapped Bose gas is expected to be altered as a direct consequence of the output coupling. In simple terms, when the output atoms are mainly those of condensates we expect the coherence of the internal atoms to drop, if only because the amount of coherent condensate fraction decreases. The coherence of trapped atoms, although interesting theoretically, is not experimentally verifiable. Apart from designing an atomic laser with well-controlled beam properties, we saw that the measured output properties may be an excellent tool in investigating the nature of trapped Bose gases at finite temperatures. The properties of the output beam may be a probe to the temperature of the trapped Bose gas as well as the internal structure of the ground state and the excitations. The present treatment may be extended to cope with other possible configurations that are likely to appear in the future such as a trap with multi-component condensates and Bose gases with negative scattering lengths. Finally we note that the pair breaking process, and indeed the output coupling of the condensate in general, provides an experimentally feasible method to study quantum entanglement in a macroscopic system. The quantum theory of entanglement is currently under intense study owing to its relevance to quantum computation; so far it has rarely been studied in the context of BEC. ###### Acknowledgements. S.C. acknowledges UK CVCP for support. Y.J. acknowledges support from the Foreign and Commonwealth Office by the British Council, from the Royal Society of London and from EC-TMR grants. This work was supported in part by the U.K. EPSRC (K.B.). ## A Expression for $`\lambda (𝐫,t)`$ – stimulated Raman scheme We derive here the expression for the effective coupling function $`\lambda (𝐫,t)`$ in Eq. (10) for the stimulated Raman transition coupling scheme. A detailed analysis of a Raman coupling process from a condensate in the mean-field approach can be found in Ref.. We consider a single atom which can be found either in the trapped level $`|t`$ or in the free level $`|f`$. A pair of laser beams with spatial and temporal amplitudes $`E_{tL}(𝐫,t)`$ and $`E_{fL}(𝐫,t)`$ are responsible for non-resonant transitions from $`|t`$ and $`|f`$ to a high energy level $`|i`$. The Hamiltonian is then given by $$H=\underset{j=t,f,i}{}[\mathrm{}\omega _j+H_j^{(0)}]|jj|+\frac{1}{2}\underset{j=t,f}{}[\mu _{ji}E_{jL}(𝐫,t)|ij|+h.c.]$$ (A1) Here $`\mathrm{}\omega _j`$ are the internal energies of the levels $`|j`$, and $$H_j^{(0)}=\frac{\mathrm{}^2^2}{2m}+V_j(𝐫),$$ (A2) where $`V_j(𝐫)`$ are effective external potentials acting on the different atomic levels, and $`\mu _{j3}`$ are the dipole moments for the transition $`|j|i`$. The amplitudes of the laser fields are assumed to have the form $`E_{jL}(𝐫,t)=_j(𝐫,t)e^{i(𝐤_{jL}𝐫\omega _{jL}t)}`$ where the envelopes $`_{jL}`$ are slowly varying with respect to the exponential term. The wave function describing the atom has the form $`_{j=t,f,i}\psi _j(𝐫,t)e^{i\omega _jt}|j`$. The equations of motion for the three amplitudes are: $`i\mathrm{}\dot{\psi _t}`$ $`=`$ $`H_t^{(0)}\psi _t\mu _{ti}^{}E_{tL}^{}\psi _ie^{i(\omega _i\omega _t)t}`$ (A3) $`i\mathrm{}\dot{\psi _f}`$ $`=`$ $`H_f^{(0)}\psi _f\mu _{fi}^{}E_{fL}^{}\psi _ie^{i(\omega _i\omega _f)t}`$ (A4) $`i\mathrm{}\dot{\psi _i}`$ $`=`$ $`H_i^{(0)}\psi _i\mu _{ti}E_{tL}\psi _te^{i(\omega _t\omega _i)t}\mu _{fL}E_{fL}\psi _fe^{i(\omega _f\omega _i)t}.`$ (A5) The solution of the equation for $`\psi _i`$ as a function of the two other amplitudes can be written as $`\psi _i(𝐫,t)`$ $`=`$ $`\psi _i^{(0)}(𝐫,t)+i{\displaystyle _0^t}𝑑t^{}{\displaystyle d^3𝐫^{}K_i(𝐫,𝐫^{},tt^{})\underset{j=t,f}{}e^{i(\omega _j\omega _i)t^{}}}`$ (A7) $`\times \mathrm{\Omega }_j(𝐫^{},t^{})\psi _j(𝐫^{},t^{})e^{i(𝐤_{jL}𝐫^{}\omega _{jL}t^{})}.`$ Here $`\psi _i^{(0)}`$ is the solution of the Schrödinger equation $`i\mathrm{}\psi _i/t=H_i^{(0)}\psi _i`$ with the initial condition $`\psi ^{(0)}(0)=\psi _i(0)=0`$, under the assumption that level $`|i`$ is initially unpopulated. $`\mathrm{\Omega }_j(𝐫,t)=\mu _{ji}_{jL}(𝐫,t)/2\mathrm{}`$ are the slowly varying Rabi frequencies and $`K_i`$ is the propagator for the evolution of the level $`|i`$, which can be expanded in terms of the energy eigenfunctions of $`H_i^{(0)}`$ in a similar way to the expansion in Eq. (23). The crucial step now is to notice that the main time-dependence in the time-integral in Eq. (A7) comes from the terms $`e^{i(\omega _j\omega _i+\omega _{jL})t}`$, whose frequency of oscillation is assumed to be in the optical range, while the other terms, which correspond to atomic centre-of-mass motion are oscillating in frequencies below the radio-frequency range. Assuming that the switching time of the coupling is much longer than the short period of oscillation of the fast terms, we expect the contribution to the integral in $`t^{}`$ to come only from a short time interval around the end-point $`t`$. We then take $`t^{}=t`$ in the slow terms. As a result, we have $`K_i(𝐫,𝐫^{},tt^{})K_i(𝐫,𝐫^{},0)=\delta (𝐫𝐫^{})`$ and $`\psi _j(𝐫^{},t^{})\psi _j(𝐫,t)`$, and obtain $$\psi _i(𝐫,t)=\underset{j=t,f}{}\frac{e^{i(\omega _j\omega _i+\omega _{jL})t}1}{\omega _j\omega _i+\omega _L}\mathrm{\Omega }_j(𝐫,t)\psi _j(𝐫,t)e^{i𝐤_{jL}𝐫}.$$ (A8) When this is substituted in Eqs. (A3), (A4), and the rapidly oscillating terms are dropped, we obtain $`\dot{\psi _t}`$ $`=`$ $`{\displaystyle \frac{i}{\mathrm{}}}H_t^{(0)}\psi _t+i\lambda _{tt}(𝐫,t)\psi _t(𝐫,t)+i\lambda _{tf}(𝐫,t)\psi _f(𝐫,t)`$ (A9) $`\dot{\psi _f}`$ $`=`$ $`{\displaystyle \frac{i}{\mathrm{}}}H_f^{(0)}\psi _f+i\lambda _{ff}(𝐫,t)\psi _f(𝐫,t)+i\lambda _{ft}(𝐫,t)\psi _t(𝐫,t),`$ (A10) where $`\lambda _{jj^{}}(𝐫,t)={\displaystyle \frac{\mathrm{\Omega }_j^{}(𝐫,t)\mathrm{\Omega }_j^{}(𝐫,t)}{\omega _j^{}\omega _i+\omega _{j^{}L}}}e^{i(\omega _{j^{}L}+\omega _j^{}\omega _{jL}\omega _j)t}e^{i(𝐤_{j^{}L}𝐤_{jL})𝐫}.`$ The form of $`\lambda (𝐫.t)`$ in Eq. (10) is achieved by assuming $`\omega _t\omega _i+\omega _{tL}\omega _f\omega _i+\omega _{fL}\mathrm{\Delta }_i`$ and then noticing that $`\lambda _{tf}=\lambda _{ft}^{}`$. We have also neglected the diagonal terms $`\lambda _{tt},\lambda _{ff}`$, which are responsible for an additional effective potential acting on the levels $`|t`$ and $`|f`$, under the assumption that they are small compared to the other potentials $`V_t(𝐫)`$ and $`V_f(𝐫)`$ near the trap. This assumption is justified in the adiabatic case discussed in this paper, where the coupling is assumed to be weak and slow. in
warning/0001/cs0001026.html
ar5iv
text
# A Logic for SDSI’s Linked Local Name Spaces ## 1 Introduction Rivest and Lampson \[RL96\] introduced SDSI—a Simple Distributed Security Infrastructure—to facilitate the construction of secure systems.<sup>1</sup><sup>1</sup>1SDSI now forms the basis for the Simple Public Key Infrastructure (SPKI) standardization work \[Gro98\]. SPKI simplifies some SDSI features (e.g., it eliminates groups) but adds many others. We focus in this paper on the core naming features of SDSI—there are some minor differences in the way that SPKI has chosen to handle these features, but we believe that our work is equally relevant to the the fragment of SPKI dealing with naming. In SDSI, principals (agents) are identified with public keys. In addition to principals, SDSI allows other names, such as poker-buddies. Rather than having a global name space, these names are interpreted locally, by each principal. That is, each principal associates with each name a set of principals. Of course, the interpretation of a name such as poker-buddies may be different for each agent. However, a principal can “export” his bindings to other principals. Thus, Ron may receive a message from the principal he names Joe describing a set of principals Joe associates with poker-buddies. Ron may then refer to the principals Joe associates with poker-buddies by the expression Joe’s poker-buddies. Rivest and Lampson \[RL96\] give an operational account of local names; they provide a name-resolution algorithm that, given a principal $`𝚔`$ and a name $`𝚗`$, computes the set of principals associated with $`𝚗`$ according to $`𝚔`$. Abadi \[Aba98\] has provided a logic that, among other things, gives a more semantic account of local names. According to Abadi, its purpose “is to explain local names in a general, self-contained way, without requiring reference to particular implementations.” Abadi shows that the SDSI name-resolution algorithm can be captured in terms of a collection of sound proof rules in his logic. Abadi’s focus is on axioms. He constructs a semantics, not with the goal of capturing the intended meaning of his constructs, but rather, with the goal of showing that certain formulas are not derivable from his axioms. (In particular, he shows that false is not derivable, showing that his axioms are consistent.) While adequate for Abadi’s restricted goals, his semantics validates some formulas that we certainly would not expect to be valid. One consequence of this is that, while he is able to pinpoint some potential concerns with the logic, the resolution of these concerns is less satisfactory. For example, he observes that adding two seemingly reasonable axioms to his logic allows us to reach quite an unreasonable conclusion. However, it is not obvious from the semantic intuitions provided by Abadi which (if either) of the axioms is unreasonable, or why it is unreasonable. Moreover, while he proves that this particular unreasonable conclusion is not derivable in his framework, as we show, a closely related (and equally unreasonable) conclusion is in fact valid. This means we have no assurance that it or other similar formulas cannot be derived from Abadi’s axioms. We very much subscribe to Abadi’s goal of using a logic to give a general account of naming. In this paper, we provide a logic whose syntax is very similar to Abadi’s, but whose semantics is quite different and, we believe, captures better the meaning we intend the constructs to have. Nevertheless, all but one of Abadi’s name space axioms are sound in our system. We remark that, in a sense, our task is much easier than Abadi’s, since we give the constructs in the logic a somewhat narrower reading than he does. Abadi tends to intertwine and occasionally identify issues of naming with issues of rights and delegation. (Such an identification is also implicitly made to some extent in designs such as PolicyMaker \[BFL96\].) We believe that it is important to treat these issues separately. Such a separation allows us to both give a cleaner semantics for each of the relevant notions and to clarify a number of subtleties. This paper focuses on naming, which we carefully separate from the other issues; a companion paper \[HvdMS99\] considers authority and delegation. We believe that our approach has a number of significant advantages: * We can still simulate the SDSI’s name resolution algorithm; Abadi’s extra axiom is unnecessary. In fact, our logic captures SDSI’s name resolution more accurately than Abadi’s. Abadi’s logic can draw conclusions that SDSI’s name resolution cannot; our logic, in a precise sense, draws exactly the same conclusions as SDSI’s name resolution algorithm. * According to our semantic intuition, one of Abadi’s proposed additional axioms is in fact quite unreasonable; it does not hold under our semantics, and it is quite clear why. * We are able to provide a sound and complete axiomatization of our logic. Thus, unlike Abadi, we have a proof system that corresponds precisely to our semantics. This will allow us to prove stronger results than Abadi’s about formulas that cannot be derived in our framework. Our completeness proof also yields a (provably optimal) NP-complete decision procedure for satisfiability of formulas in the logic. * Our logic is closely related to Logic Programming. This allows us to translate queries about names to Logic Programming queries, and thus use all the well-developed Logic Programming technology to deal with such queries. * Our approach opens the road to a number of generalizations, which allow us to deal with issues like permission, authority, and delegation \[HvdMS99\]. The rest of this paper is organized as follows. In Section 2, we review Abadi’s logic and, in the process, describe SDSI’s naming scheme. We also point out what we see as the problems with Abadi’s approach. In Section 3, we give the syntax and semantics of our logic, and present a complete axiomatization. In Section 4 we show that our logic provides a tight characterization of SDSI name resolution. Section 5 deals with the connection between our account of SDSI name resolution and logic programming, and Section 6 concerns $`\mathrm{𝚂𝚎𝚕𝚏}`$, an additional construct considered by Abadi. Section 7 concludes. ## 2 SDSI’s Name Spaces and Abadi’s Logic In this section, we briefly review SDSI’s naming scheme and Abadi’s logic, and discuss our criticism of Abadi’s logic. Like Abadi, we are basing our discussion on SDSI 1.1 \[RL96\]. ### 2.1 SDSI’s Name Spaces SDSI has local names and a set of reserved names, which we refer to as global names. Both are associated with sets of principals, but the set of principals associated with a local name depends on the principal owning the local name space, while the set of principals associated with a global name does not. We denote the set of global names by $`G`$ with generic element $`𝚐`$, the set of local names by $`N`$ with generic element $`𝚗`$, and the set of keys (principals) by $`K`$ with generic element $`𝚔`$. We assume that all these sets are pairwise disjoint and that $`K`$ is nonempty. Global identifiers are either keys or global names.<sup>2</sup><sup>2</sup>2Note that Abadi uses $`G`$ for global identifier; thus, his $`G`$ corresponds to our $`GK`$. The elements of $`KGN`$ are said to be simple names. We form principal expressions from simple names inductively. Simple names are principal expressions, and if $`𝚙`$ and $`𝚚`$ are principal expressions, then so is $`(𝚙\text{’s}𝚚)`$. Abadi’s semantics (and ours) makes the latter operation associative, in that $`((𝚙\text{’s}𝚚)\text{’s}𝚛)`$ and $`(𝚙\text{’s}(𝚚\text{’s}𝚛))`$ have the same meaning. In light of this, we can ignore parenthesization when writing such expressions. The expression $`𝚙_1`$’s …$`𝚙_{m1}`$’s $`𝚙_m`$ is written in SDSI as $`(\mathrm{𝚛𝚎𝚏}:𝚙_1,\mathrm{},𝚙_m)`$.<sup>3</sup><sup>3</sup>3SDSI allows $`m`$ to be 0, taking (ref:) to be the current principal. In Section 6, we follow Abadi by considering an expression $`\mathrm{𝚂𝚎𝚕𝚏}`$ that represents (ref:). We remark for future reference that SDSI has a special global name denoted “$`\mathrm{𝙳𝙽𝚂}!!`$”, which represents the root of the DNS (Internet mail) hierarchy; this allows us to express an email address such as bob@fudge.com as $`\mathrm{𝙳𝙽𝚂}!!`$’s com’s fudge’s bob. SDSI allows a principal to issue certificates of the form $`𝚗𝚙`$, signed with its key. If $`𝚔`$ issues such a certificate, it has the effect of binding local name $`𝚗`$ in $`𝚔`$’s name space to the principals denoted by the principal expression $`𝚙`$.<sup>4</sup><sup>4</sup>4SDSI also allows other forms of binding that we do not consider here. Our notation is also a simplification of that used by SDSI. Notice that only principals issue certificates, and that these certificates bind a local name (not a global name) to some set of principals. In general, a local name may be bound to a unique principal, no principal, or many principals. SDSI allows a principal $`𝚔`$ to issue certificates $`𝚗𝚙_1`$ and $`𝚗𝚙_2`$. This has the effect of binding $`𝚗`$ to (at least) the principals denoted by $`𝚙_1`$ and $`𝚙_2`$. SDSI provides a name-resolution algorithm for computing the set of principals bound to a name. The core of the algorithm consists of a nondeterministic procedure REF2. For ease of exposition, we take REF2 to have four arguments: a principal $`𝚔`$, a function $`c`$ that associates with each principal $`𝚔^{}`$ a set of bindings (intuitively, ones that correspond to certificates signed by $`𝚔^{}`$), a function $`\beta `$ which associates with each global name $`𝚐`$ a set of principals (intuitively, the ones bound to $`𝚐`$), and a principal expression $`𝚙`$. REF2($`𝚔`$,$`\beta `$,$`c`$,$`𝚙`$) returns the principal(s) bound to $`𝚙`$ in $`𝚔`$’s name space, given the bindings $`\beta `$ and the certificates $`c`$. REF2 is nondeterministic; the set of possible outputs of REF2 is taken to be the set of principals bound to $`𝚙`$ in $`𝚔`$’s name space. REF2 is described in Figure 1.<sup>5</sup><sup>5</sup>5Our version of REF2 is similar, although not identical, to Abadi’s. Like Abadi’s, it is simpler than that in \[RL96\], in that we do not deal with a number of issues, such as quoting or encrypted objects, dealt with by SDSI. Our presentation of REF2 differs from Abadi’s mainly in its treatment of global names. Abadi assumes that REF2 takes only two arguments, o and p, where o is either a global identifier (i.e., an element of $`GK`$) or current principal, denoted cp. Although he does not write $`c`$ explicitly as an argument, he does assume that there is a set he denotes assumptions(o) that includes bindings corresponding to signed certificates. In addition, it includes bindings for cp. We do not have a distinguished current principal; rather, if the current principal is $`𝚔`$, then for uniformity we assume that all of the current principal’s bindings are also described by the bindings in $`c(𝚔)`$. More significantly, if g is a global name, then Abadi’s REF2(o,g) would return g, while ours would return some principal k to which g is bound in $`\beta `$. Our approach seems more consistent with the SDSI presentation of REF2, but this difference is minor, and all of Abadi’s results hold for our presentation of REF2. ### 2.2 Abadi’s Logic: Syntax, Semantics, and Axiomatization The formulas in Abadi’s logic are formed by starting with a set of primitive propositions and formulas of the form $`𝚙𝚙^{}`$, where $`𝚙`$ and $`𝚙^{}`$ are principal expressions. More complicated formulas are formed by closing off under conjunction, negation, and formulas of the form $`𝚙\mathrm{𝑠𝑎𝑦𝑠}\varphi `$, where $`\varphi `$ is a formula. Abadi views $`𝚙𝚙^{}`$ as meaning that $`𝚙`$ is “bound to” $`𝚙^{}`$. He considers two possible interpretations of “bound to”. The first is equality; however, he rejects this as being inappropriate. (In particular, it does not satisfy some of his axioms.) The second is that $`𝚙𝚙^{}`$ means $`𝚙^{}`$ “speaks-for” $`𝚙`$, in the sense discussed in \[ABLP93, LABW92\]. Roughly speaking, this says that any message certified by $`𝚙^{}`$ should be viewed as also having been certified by $`𝚙`$. While the “speaking-for” interpretation is the one favored by Abadi, he does not commit to it. Note that under Abadi’s “speaking-for” interpretation, it makes sense to write $`𝚙𝚙^{}`$ for arbitrary principal expressions $`𝚙`$ and $`𝚙^{}`$. However, SDSI allows only local (simple) names to be bound to principal expressions. We shall make a similar restriction in our logic (and, indeed, under our semantic interpretation of binding, it would not make sense to allow an arbitrary principal expression to be bound to another one.) The “speaks-for” interpretation intertwines issues of delegation with those of naming. As we suggested in the introduction, we believe these issues should be separated. We shall give $``$ a different interpretation that we believe is simpler and more in the spirit of binding. We believe that the “speaks-for” relation of \[ABLP93, LABW92\] should have quite different semantics than that of binding names to principals. (We hope to return to this issue in future work.) Abadi interprets $`𝚙\mathrm{𝑠𝑎𝑦𝑠}\varphi `$ as “the principal denoted by $`p`$ makes a statement that implies $`\varphi `$”. In the case where $`𝚙`$ is a key (i.e., principal) $`𝚔`$, this could mean that $`𝚔`$ signs a statement saying $`\varphi `$. Under our more restrictive interpretation, this is exactly how we interpret our analogue to $`\mathrm{𝑠𝑎𝑦𝑠}`$. In any case, note that Abadi translates SDSI’s local name $`𝚗`$ being bound to $`𝚙`$ as $`𝚗𝚙`$ and captures $`𝚔`$ signing a certificate saying $`𝚗`$ is bound to $`𝚙`$ by the formula $`𝚔\mathrm{𝑠𝑎𝑦𝑠}𝚗𝚙`$. For future reference, it is worth noting that, in order to capture the binding of names to principals, no use is made of primitive propositions. Abadi interprets formulas in his logic with respect to a tuple $`(𝒲,\alpha ,\rho ,\mu )`$. The function $`\alpha `$ maps global identifiers ($`GK`$) to subsets of $`𝒲`$. The function $`\rho `$ maps $`N\times 𝒲`$ to subsets of $`𝒲`$. Finally, $`\mu `$ associates with each world (principal) $`𝚔`$ and primitive proposition $`p`$ a truth value $`\mu (p,𝚔)`$. Abadi does not provide any intuition for his semantics, but suggests that $`𝒲`$ should be thought of as a set of possible worlds, as in modal logic. However, he also suggests \[private communication, 1999\] that his semantics was motivated by the work of Grove and Halpern \[GH93\], in which the corresponding set contains pairs consisting of a world and an agent. Some of Abadi’s definitions make more intuitive sense if we think of $`𝒲`$ as a set of agents, while others make more sense if we think of $`𝒲`$ as a set of worlds. We elaborate on this point below. Given $`𝚔𝒲`$ and $`𝚙P`$, Abadi defines $`[[𝚙]]_𝚔`$ inductively, as follows: * $`[[𝚐]]_𝚔=\alpha (𝚐)`$, for $`𝚐GK`$ * $`[[𝚗]]_𝚔=\rho (𝚗,𝚔)`$ for $`𝚗N`$ * $`[[𝚙_1\text{’s }𝚙_2]]_𝚔=\{[[𝚙_2]]_𝚔^{}:𝚔^{}[[𝚙_1]]_𝚔\}`$ Here we have used a notation corresponding to the interpretation of the “worlds” in $`𝒲`$ as agents. Using this interpretation we may think of $`[[𝚙]]_𝚔`$ as the set of principals bound to principal expression $`𝚙`$ according to $`𝚔`$. The clause for $`[[𝚙_1\text{’s }𝚙_2]]_𝚔`$ then says that if $`𝚔^{}`$ is one of the principals referred to by $`𝚔`$ as $`𝚙_1`$, then $`𝚔`$ uses $`𝚙_1\text{’s}𝚙_2`$ to refer to any principal referred to by $`𝚔^{}`$ as $`𝚙_2`$. Abadi also defines what it means for a formula $`\varphi `$ to be true at world $`𝚔𝒲`$, written $`𝚔\varphi `$, inductively, by * $`𝚔p`$ iff $`\mu (p,𝚔)=\mathrm{𝐭𝐫𝐮𝐞}`$, if $`p`$ is a primitive proposition * $`𝚔\varphi \psi `$ iff $`𝚔\varphi `$ and $`𝚔\psi `$ * $`𝚔\neg \varphi `$ iff $`𝚔\vDash ̸\varphi `$ * $`𝚔𝚙𝚙^{}`$ iff $`[[𝚙]]_𝚔[[𝚙^{}]]_𝚔`$ * $`𝚔𝚙\mathrm{𝑠𝑎𝑦𝑠}\varphi `$ iff $`𝚔^{}\varphi `$ for all $`𝚔^{}[[𝚙]]_𝚔`$. These clauses defining $``$ are quite intuitive if one interprets $`𝒲`$ to be a set of worlds and considers $`[[𝚙]]_𝚔`$ to be the set of worlds consistent with what principal $`𝚙`$ has said at world $`𝚔`$. In particular, under this interpretation, the clause for $`\mathrm{𝑠𝑎𝑦𝑠}`$ can be read as stating that $`𝚙\mathrm{𝑠𝑎𝑦𝑠}\varphi `$ if $`\varphi `$ holds in all worlds consistent with what $`𝚙`$ has said. The clause for $``$ also has quite a plausible reading under the “speaks-for” interpretation of this construct: it states that $`𝚙^{}`$ speaks for $`𝚙`$ if all worlds consistent with what $`𝚙`$ has said are consistent with what $`𝚙^{}`$ has said, i.e., $`𝚙`$ is constrained to speak consistently with what $`𝚙^{}`$ has said. However, it seems rather difficult to extend this intuitive reading to encompass the inductive definition of $`[[𝚙]]_𝚔`$. In particular, it is far from clear to us what intuitive understanding to assign to the clause for $`[[𝚙_1\text{’s }𝚙_2]]_𝚔`$ on this reading. On the other hand, note that if we interpret the worlds as agents, then we can think of $`𝚔\varphi `$ as saying that $`\varphi `$ is true when local names are interpreted according to agent $`𝚔`$. But this reading of the clauses, when combined with the intuitive reading of $`[[𝚙]]_𝚔`$ as the set of principals that $`𝚔`$ refers to using $`𝚙`$, also has its difficulties. Intuitively, when $`𝚗`$ is bound to $`𝚙`$ in principal $`𝚔`$’s local name space, the principals that $`𝚔`$ refers to using $`𝚙`$ should be a subset of the principals that $`𝚔`$ refers to using $`𝚗`$. Abadi interprets $`𝚗`$ being bound to $`𝚙`$ as $`𝚗𝚙`$; this holds with respect to principal $`𝚔`$ when $`[[𝚙]]_𝚔`$ is a superset of $`[[𝚗]]_𝚔`$. This is precisely the opposite of what we would expect. Thus, neither the interpretation of $`𝒲`$ as a set of worlds nor the interpretation of $`𝒲`$ as a set of agents gives a fully satisfactory justification for Abadi’s semantics. As we shall see, in our semantics, the interpretation of a principal expression $`𝚙`$ according to an agent will be a set of agents, but we use the reverse of Abadi’s containment to represent binding. Abadi provides an axiom system for his logic, which has three components: 1. The standard axioms and rules of propositional logic. 2. The standard axiom and rule for modal logic for the $`\mathrm{𝑠𝑎𝑦𝑠}`$ operator: $$(𝚙\mathrm{𝑠𝑎𝑦𝑠}(\varphi \psi ))((𝚙\mathrm{𝑠𝑎𝑦𝑠}\varphi )(𝚙\mathrm{𝑠𝑎𝑦𝑠}\psi ))$$ $$\begin{array}{c}\varphi \\ \\ 𝚙\mathrm{𝑠𝑎𝑦𝑠}\varphi \end{array}$$ 3. New axioms dealing with linked local name spaces, shown in Figure 2. He shows that this axiomatization is sound, but conjectures it is not complete. ### 2.3 Name Resolution in Abadi’s Logic Abadi proves a number of interesting results relating his logic to SDSI. First, he shows that in a precise sense his logic can simulate REF2. He provides a collection of name-resolution rules NR and proves the following results:<sup>6</sup><sup>6</sup>6The results stated here are a variant of those stated in Abadi’s paper, since our version of REF2 differs slightly from his. Nevertheless, the proofs of the results are essentially identical. ###### Proposition 2.1 : Given a collection of $`c`$ of bindings corresponding to signed certificates and a set $`\beta `$ of bindings of global names to keys, let $`E`$ be the conjunction of the formulas $`𝚔\mathrm{𝑠𝑎𝑦𝑠}𝚗𝚚`$ for each certificate $`𝚗𝚚c(𝚔)`$ and the formulas $`𝚐𝚔`$ for each $`𝚔\beta (𝚐)`$. Then $`E((𝚔\text{’s}𝚙)𝚔_1)`$ is provable with the name resolution rules NR if and only if $`\text{REF2}(𝚔,\beta ,c,𝚙)`$ yields $`𝚔_1`$. ###### Proposition 2.2 : The name resolution rules are sound with respect to the logic. That is, given $`E`$ as in Proposition 2.1 and any principal expression $`𝚙`$, if $`E(𝚙𝚔)`$ is provable using NR then $`E(𝚙𝚔)`$ is also provable in the logic. These results show that any bindings of names to principals that can be deduced using REF2 can also be deduced using Abadi’s logic. However, Abadi shows that his logic is actually more powerful than REF2, by giving two examples of conclusions that can be deduced from his logic but not using REF2: ###### Example 2.3 : Using the Globality, Associativity, and Transitivity axioms, if $`𝚔`$ and $`𝚔^{}`$ are keys, we immediately get $`𝚔\text{’s}(\mathrm{𝙻𝚊𝚖𝚙𝚜𝚘𝚗}\text{’s}𝚔^{})𝚔^{}`$. This result does not follow from the REF2 algorithm. That is, $`\text{REF2}(𝚔,\beta ,c,\mathrm{𝙻𝚊𝚖𝚙𝚜𝚘𝚗}\text{’s}𝚔^{}`$) does not necessarily yield $`𝚔^{}`$ for arbitrary $`c`$ and $`\beta `$ (in particular, it will not do so if Lampson is not bound to anything in $`c`$). ###### Example 2.4 : Suppose $`c`$ consists of the four certificates that correspond to the following formulas: $`𝚔\mathrm{𝑠𝑎𝑦𝑠}(\mathrm{𝙻𝚊𝚖𝚙𝚜𝚘𝚗}𝚔_1)`$, $`𝚔\mathrm{𝑠𝑎𝑦𝑠}(\mathrm{𝙻𝚊𝚖𝚙𝚜𝚘𝚗}𝚔_2)`$, $`𝚔_1\mathrm{𝑠𝑎𝑦𝑠}(\mathrm{𝚁𝚘𝚗}\mathrm{𝚁𝚒𝚟𝚎𝚜𝚝})`$, and $`𝚔_2\mathrm{𝑠𝑎𝑦𝑠}(\mathrm{𝚁𝚒𝚟𝚎𝚜𝚝}𝚔_3)`$ (where $`𝚔`$, $`𝚔_1`$, $`𝚔_2`$, and $`𝚔_3`$ are keys). Using the Speaking-for axiom, it is not hard to show that we can conclude that $`𝚔\text{’s}(\text{Lampson’s Ron})𝚔_3`$. It is easy to show that REF2 cannot reach this conclusion; that is, $`\text{REF2}(𝚔,\beta ,c,\text{Lampson’s Ron})`$ does not yield $`𝚔_3`$ for any $`\beta `$.<sup>7</sup><sup>7</sup>7SPKI certificates and SDSI certificates have a slightly different syntactic form. A SPKI certificate issued by $`𝚔`$ to bind $`𝚗`$ to $`𝚙`$ could be expressed in the logic as $`𝚔\mathrm{𝑠𝑎𝑦𝑠}(𝚔\text{’s}𝚗𝚙)`$. Abadi has remarked \[private communication 1999\], that if we rewrite the example using assertions in this form, the corresponding conclusion of this example would not follow in his logic. We have followed the SDSI format for certificates in this paper, but note that after some minor changes to the definitions, all the results in Sections 3–5 would still apply to SPKI certificates. In reference to Example 2.3, Abadi \[Aba98\] says that “it is not clear whether \[these conclusions\] are harmful, and they might in fact be useful”. In general, he views it as a feature of his logic that it allows reasoning about names without knowing their bindings \[private communication, 1999\]. While we agree that, in general, reasoning about names without knowing their bindings is a powerful feature, we believe it is important to make clear exactly which conclusions are desirable and which are not. This is what a good semantics can provide. Under our semantics, neither of these two conclusions are valid. In fact, our logic draws precisely the same conclusions as REF2. Of course, the conclusions of Examples 2.3 and 2.4 are valid under Abadi’s semantics but, as we observed earlier, Abadi’s semantics is not really meant to be used as a guide to which conclusions are acceptable (and, indeed, as we shall see, it validates a number of conclusions that do not seem so acceptable). Abadi also considers the effect of extending his axiom system. In particular, he considers adding the following two axioms: * the converse of Globality: $`𝚐(𝚙\text{’s}𝚐)`$ * a generalization of Linking: $`(𝚙\mathrm{𝑠𝑎𝑦𝑠}(𝚙_1𝚙_2))(𝚙\text{’s}𝚙_1𝚙\text{’s}𝚙_2)`$, for an arbitrary principal $`𝚙_1`$ (instead of a local name). The generalization of Linking is in fact sound under Abadi’s semantics. The converse of Globality is not, but only because we may have $`[[𝚙]]_𝚔=\mathrm{}`$. Note that $`[[𝚙]]_𝚔=\mathrm{}`$ iff $`𝚔𝚙\mathrm{𝑠𝑎𝑦𝑠}\text{false}`$; thus, the following variant of the converse of Globability is sound under Abadi’s semantics: $`\neg (𝚙\mathrm{𝑠𝑎𝑦𝑠}\text{false})(𝚐(𝚙\text{’s}𝚐))`$. This is quite relevant to our purposes because Abadi shows that if we added the two axioms above to his system, then from $`𝚔\mathrm{𝑠𝑎𝑦𝑠}(\mathrm{𝙳𝙽𝚂}!!𝚔)`$, we can conclude $`\mathrm{𝙳𝙽𝚂}!!𝚔`$. Thus, just from $`𝚔`$ saying that $`\mathrm{𝙳𝙽𝚂}!!`$ is bound to $`𝚔`$, it follows that $`\mathrm{𝙳𝙽𝚂}!!`$ is indeed bound to $`𝚔`$. This is particularly disconcerting under Abadi’s “speaks-for” interpretation, where $`\mathrm{𝙳𝙽𝚂}!!𝚔`$ becomes “$`𝚔`$ speaks for DNS!!”. We certainly do not want an arbitrary principal to speak for the name server! Abadi proves a result showing that such conclusions are not derivable from hypotheses of a certain type in his logic (which does not have these two axioms). ###### Proposition 2.5 : \[Aba98\] Let $`𝚔`$ and $`𝚔^{}`$ be distinct global names; let $`\varphi `$ be a formula of the form $`(𝚔^{}\mathrm{𝑠𝑎𝑦𝑠}(𝚗_1𝚙_1))\mathrm{}(𝚔^{}\mathrm{𝑠𝑎𝑦𝑠}(𝚗_k𝚙_k))`$, where $`𝚗_1,\mathrm{},𝚗_k`$ are local names and $`𝚙_1,\mathrm{},𝚙_k`$ are principal expressions; let $`\psi `$ be a formula of the form $`(𝚔\mathrm{𝑠𝑎𝑦𝑠}\psi _1)\mathrm{}(𝚔\mathrm{𝑠𝑎𝑦𝑠}\psi _m)`$, where $`\psi _1,\mathrm{},\psi _m`$ are arbitrary formulas. Then $`\varphi \psi (𝚔^{}𝚔)`$ is not valid.<sup>8</sup><sup>8</sup>8Abadi’s result actually says “$`\varphi \psi (𝚔^{}𝚔)`$ is not derivable”; since his axiomatization is sound, but not necessarily complete, the claim that it is not valid is stronger, and that is what Abadi’s proof shows. While Proposition 2.5 provides some assurance that undesirable formulas are not derivable in the logic, it does not provide much. Indeed, if we allow the $`\psi `$ to include the formula $`\neg (𝚔^{}`$ $`\mathrm{𝑠𝑎𝑦𝑠}\text{false})`$, then the result no longer holds. In fact, it follows from our earlier discussion that the formula $$(𝚔\mathrm{𝑠𝑎𝑦𝑠}(\mathrm{𝙳𝙽𝚂}!!𝚔))\neg (𝚔\mathrm{𝑠𝑎𝑦𝑠}\text{false})(\mathrm{𝙳𝙽𝚂}!!𝚔)$$ is valid. Moreover, it does not seem so unreasonable to allow conjuncts such as $`\neg (𝚔\text{ }\mathrm{𝑠𝑎𝑦𝑠}\text{ }\text{false})`$ as part of $`\psi `$. We certainly want to be able to use the logic to be able to say that if a principal’s statements are not blatantly inconsistent, then certain conclusions follow. ## 3 The Logic of Local Name Containment In this section we propose the Logic of Local Name Containment (henceforth LLNC) as an alternative to Abadi’s logic. LLNC interprets local names as sets of principals and interprets SDSI certificates as stating containment relationships between these sets. We define the syntax in Section 3.1. In Section 3.2 we describe two distinct semantics for the logic. Section 3.3 presents a complete axiomatization. ### 3.1 Syntax LLNC has syntactic elements that are closely related to the syntactic elements of Abadi’s logic. However, our notation differs slightly from Abadi’s to help emphasize some of the differences in intuition. Again, we start with keys $`K`$, global names $`G`$, and local names $`N`$, and form principal expressions from them. The formulas of our language are formed as follows: * If $`𝚙`$ and $`𝚚`$ are principal expressions then $`𝚙𝚚`$ is a formula. * If $`𝚔K`$ and $`\varphi `$ is a formula then $`𝚔\mathrm{𝑐𝑒𝑟𝑡}\varphi `$ is a formula.<sup>9</sup><sup>9</sup>9For our account of SDSI naming, it would suffice to restrict this clause to formulas of the form $`𝚔\mathrm{𝑐𝑒𝑟𝑡}𝚗𝚙`$ where $`𝚗N`$ and $`𝚙P`$: our semantics will treat more general certificates as irrelevant to the meaning of principal expressions. We allow the more general form for purposes of discussion and because we envisage generalizations of the logic in which other types of certificates will be required. * If $`\varphi _1`$ and $`\varphi _2`$ are formulas, then so are $`\neg \varphi _1`$ and $`\varphi _1\varphi _2`$. As usual, $`\varphi _1\varphi _2`$ is an abbreviation for $`\neg (\neg \varphi _1\neg \varphi _2)`$ and $`\varphi _1\varphi _2`$ is an abbreviation for $`\neg \varphi _1\varphi _2`$. We write $``$ for the set of all formulas. (For simplicity, we omit primitive propositions, although we could easily add them. They play no role in Abadi’s account of SDSI names, nor will they in ours.) We read the expression $`𝚙𝚚`$ as “$`𝚙`$ contains $`𝚚`$”; we intend for it to capture the fact that all the keys bound to $`𝚚`$ are also bound to $`𝚙`$. However, our intuitions about the meaning of $`𝚙𝚚`$ are quite different from Abadi’s. In particular, we do not wish to interpret $`𝚙𝚚`$ as “$`𝚚`$ speaks for $`𝚙`$.” We consider the “speaks for” relation as being about rights and delegation, which requires a more sophisticated semantics than we wish to consider here. (See \[HvdMS99\] for a logic for reasoning about rights and delegation.) The expression $`𝚙𝚚`$ should be understood as simply asserting a containment relationship between the denotations of principal expressions $`𝚙`$ and $`𝚚`$; this is exactly what our semantics will enforce. We read the expression $`𝚔\mathrm{𝑐𝑒𝑟𝑡}\varphi `$ as “$`𝚔`$ has certified that $`\varphi `$.” This corresponds roughly to Abadi’s $`𝚔\mathrm{𝑠𝑎𝑦𝑠}\varphi `$. There are two significant differences, however. For one thing, we do not allow arbritrary principal expressions on the left-hand side; only keys may certify a formula $`\varphi `$. For another, our interpretation of $`\mathrm{𝑐𝑒𝑟𝑡}`$ is more restrictive than Abadi’s $`\mathrm{𝑠𝑎𝑦𝑠}`$, in that $`\mathrm{𝑐𝑒𝑟𝑡}`$ is treated quite syntactically; it refers to an actual certificate issued by a principal, while $`\mathrm{𝑠𝑎𝑦𝑠}`$ considers logical consequences of such certificates. As a consequence, whereas $`\mathrm{𝑠𝑎𝑦𝑠}`$ satisfies standard properties of modal operators (e.g., closure under logical consequence), $`\mathrm{𝑐𝑒𝑟𝑡}`$ does not. ### 3.2 Semantics Our semantics is designed to model the SDSI principle that principals bind names in their local name space to values by issuing certificates. The interpretation of a local name depends on the principal and the certificates that have been issued. As the principal may rely on others for its interpretation of local names, the certificates issued by other principals also play a role. The interpretation of global names and keys will be independent of both the principal and the certificates that have been issued. A world is a pair $`w=(\beta ,c)`$, where $`\beta :G𝒫(K)`$ and $`c:K𝒫()`$ (where $`𝒫(X)`$ denotes the set of subsets of $`X`$) and $`_{𝚔K}c(𝚔)`$ is finite. Intuitively, the function $`\beta `$ interprets global (or fixed) names as sets of keys. The intended interpretation of the function $`c`$ is that it associates with every key $`𝚔`$ the set of formulas $`c(𝚔)`$ that have been certified using this key. That is, if $`\varphi c(𝚔)`$ then, intuitively, a certificate asserting $`\varphi `$ has been signed using $`𝚔`$.<sup>10</sup><sup>10</sup>10We make the simplifying assumption that certificates do not have expiration dates. It is not difficult to extend the logic to take into account certificate expiration; see \[HvdM99\]. The assumption that $`_{𝚔K}c(𝚔)`$ is finite is meant to enforce the intuition that only finitely many certificates are issued. None of our later results depend on this assumption, but it seems reasonable given the intended application of the logic. Formulas of the logic will be interpreted in a world with respect to a key. Intuitively, this key indicates the principal from whose perspective we interpret principal expressions. To interpret local names, we introduce an additional semantic construct. A local name assignment will be a function $`l:K\times N𝒫(K)`$ associating each key and local name with a set of keys. Intuitively, $`l(𝚔,𝚗)`$ is the set of keys represented by principal $`𝚔`$’s local name $`𝚗`$. We write $`\mathrm{𝐿𝑁𝐴}`$ for the set of all local name assignments. Given a world $`w=(\beta ,c)`$, a local name assignment $`l`$, and a key $`𝚔`$, we may assign to each principal expression $`𝚙`$ an interpretation $`[[𝚙]]_{w,l,𝚔}`$, a set of keys. The definition is much like that of Abadi’s $`[[p]]_𝚔`$: * $`[[𝚔^{}]]_{w,l,𝚔}=\{𝚔^{}\}`$, if $`𝚔^{}K`$ is a key, * $`[[𝚐]]_{w,l,𝚔}=\beta (𝚐)`$, if $`𝚐G`$ is a global name, * $`[[𝚗]]_{w,l,𝚔}=l(𝚔,𝚗)`$, if $`𝚗N`$ is a local name, * $`[[𝚙\text{’s}𝚚]]_{w,l,𝚔}=\{[[𝚚]]_{w,l,𝚔^{}}|𝚔^{}[[𝚙]]_{w,l,𝚔}\}`$, for principal expressions $`𝚙,𝚚P`$. Our intuitions for $`[[𝚙]]_{w,l,𝚔}`$ are essentially the same as for the “agent-based” reading of Abadi’s logic, discussed above. That is, $`[[𝚙]]_{w,l,𝚔}`$ is the set of keys associated with the expression $`𝚙`$ in $`𝚔`$’s local name space, when local names are interpreted according to $`l`$. With respect to principal $`𝚔`$, the expression $`𝚙\text{’s}𝚚`$ denotes the set of principals that principals referred to by $`𝚔`$ as $`𝚙`$ refer to as $`𝚚`$. We now define what it means for a formula $`\varphi `$ to be true at a world $`w=(\beta ,c)`$ with respect to a local name assignment $`l`$ and key $`𝚔`$, written $`w,l,𝚔\varphi `$, by induction on the structure of $`\varphi `$.<sup>11</sup><sup>11</sup>11Note that our semantics is thus in the spirit of that of Grove and Halpern \[GH93\], in that the truth of a formula depends on both an agent and some features of the world (captured by $`w`$ and $`l`$). * $`w,l,𝚔𝚙𝚚`$ if $`[[𝚙]]_{w,l,𝚔}[[𝚚]]_{w,l,𝚔}`$ * $`w,l,𝚔𝚔^{}\mathrm{𝑐𝑒𝑟𝑡}\varphi `$ if $`\varphi c(𝚔^{})`$ * $`w,l,𝚔\neg \varphi _1`$ if not $`w,l,𝚔\varphi _1`$ * $`w,l,𝚔\varphi _1\varphi _2`$ if $`w,l,𝚔\varphi _1`$ and $`w,l,𝚔\varphi _2`$. Note that the semantics of $`\mathrm{𝑐𝑒𝑟𝑡}`$ reinforces its syntactic nature. To determine if $`𝚔^{}\mathrm{𝑐𝑒𝑟𝑡}\varphi `$ is true at $`(w,l,𝚔)`$, we check whether a certificate has been issued in world $`w`$ by $`𝚔^{}`$ certifying $`\varphi `$. Moreover, as we shall see, while we allow any formula to be certified by $`𝚔`$, the only formulas whose certification has a nontrivial semantic impact are those of the form $`𝚗𝚙`$, where $`𝚗`$ is a local name. We return to this issue below. We do not consider all pairs $`w,l`$ as being appropriate on the left-hand side of $``$. If $`w=(\beta ,c)`$, we expect the local name assignment $`l`$ to respect the certificates that have been issued in $`c`$. That is, if $`c(𝚔)`$ includes the binding $`𝚗𝚙`$, we would expect that $`l(𝚔,𝚗)`$ would include all the keys bound to $`𝚙`$ in $`𝚔`$’s name space. The question is whether there can be other keys bound to $`𝚗`$ in $`𝚔`$’s name space beyond those forced by the certificates. How we answer this question depends on our intuitions for $`c`$. For example, we could view $`c`$ as the set of certificates received by one of the principals. This would be particularly appropriate if we wanted to reason about the knowledge and belief of the agents, an extension we plan to explore in future work. With this viewpoint, we could view $`l`$ as consisting of all the bindings, including ones that the principal does not know about. Thus, $`l`$ would at least have all the bindings forced by $`c`$, but perhaps others as well. Alternatively, we could view $`c`$ as consisting of all the certificates that have been issued. In this case, we would want $`l`$ to be in some sense minimal, and have no bindings beyond those forced by the certificates in $`c`$. We now present two different semantics, which reflect each of these two intuitions. We then show that, as far as validity is concerned, the semantics are equivalent; that is, they have the same proof theory. A local name assignment $`l`$ is consistent with a world $`w=(\beta ,c)`$ if, for all keys $`𝚔`$, local names $`𝚗`$, and principal expressions $`𝚙`$, if the formula $`𝚗𝚙`$ is in $`c(𝚔)`$, then $`w,l,𝚔𝚗𝚙`$. Intuitively, assignments that are not consistent with a world provide an inappropriate basis for the interpretation of local names, since the certificates issued by principals are not necessarily reflected in their local bindings. We obtain our first semantics, called the open semantics, by restricting to consistent local name assignments. We write $`w,l,𝚔_\mathrm{o}\varphi `$ if $`w,l,𝚔\varphi `$ and $`l`$ is consistent with $`w`$. The formula $`\varphi `$ is o-satisfiable if there exists a triple $`w,l,𝚔`$ such that $`w,l,𝚔_\mathrm{o}\varphi `$ and $`\varphi `$ is o-valid, denoted $`_\mathrm{o}\varphi `$, if there does not exist a triple $`w,l,𝚔`$ such that $`w,l,𝚔_\mathrm{o}\neg \varphi `$. Although our syntax allows $`𝚔`$ to certify arbitrary formulas, it is easy to see that, according to the semantics just introduced (as well as the one we are about to introduce), only the certification of formulas of the form $`𝚗𝚙`$ has any impact on consistency; all other formulas certified by $`𝚔`$ are ignored. There is a good reason for this restriction. We are implicitly assuming that when $`𝚔^{}`$ certifies $`𝚗𝚙`$, that very act causes all the keys bound to $`𝚙`$ to also be bound to $`𝚗`$ in $`𝚔`$’s name space. Thus, if $`𝚗𝚙c(𝚔)`$, then we want $`𝚗𝚙`$ to be true in $`(w,l,𝚔)`$. But if $`𝚔`$ certifies a formula like $`𝚔_1`$’s $`𝚗𝚔_3`$ where $`𝚔_1𝚔`$, then we cannot conclude that this formula is true in $`(w,l,𝚔)`$ unless we are prepared to make additional assumptions about $`𝚔`$’s truthfulness. We feel that if such assumptions are to be made, then they should be modeled explicitly in the logic, not hidden in the semantics. It does seem reasonable to extend the notion of $`l`$ being consistent with $`w`$ to require that if $`𝚔`$ certifies a formula $`\psi `$ which is a Boolean combination of formulas of the form $`𝚗𝚙`$ then $`(w,l,𝚔)\psi `$. However, once we allow more general Boolean combinations (in particular, once we allow disjunctions), there will be problems making sense out of the intuition of our next semantics, that there are “no bindings beyond those forced by the certificates in $`c`$”. We consider this issue next. According to the open semantics, it is possible for a local name $`𝚗`$ of principal $`𝚔_1`$ to be bound to a key $`𝚔_2`$ even when no certificate concerning $`𝚗`$ has been issued. Arguably, this is not in accordance with the intentions of SDSI. To better capture these intentions, we define a second semantics, that restricts the name bindings to those forced by the certificates issued. To do so, we first establish that the open semantics satisfies a kind of “minimal model” result. Define the ordering $``$ on the space $`\mathrm{𝐿𝑁𝐴}`$ of local name assignments by $`l_1l_2`$ if $`l_1(𝚔,𝚗)l_2(𝚔,𝚗)`$ for all $`𝚔K`$ and $`𝚗N`$. It is readily seen that $`\mathrm{𝐿𝑁𝐴}`$ is given the structure of a complete lattice \[Bir67\] by this relation. Say that a local name assignment $`l`$ is minimal in a set of local name assignments $`L`$ if $`lL`$ and $`ll^{}`$ for all $`l^{}L`$. ###### Theorem 3.1 : Given a world $`w`$, there exists a unique local name assignment $`l_w`$ minimal in the set of all local name assignments consistent with $`w`$. Moreover, if $`𝚙`$ is a principal expression and $`𝚔_1`$ and $`𝚔_2`$ are keys, then $`w,l_w,𝚔_1_\mathrm{o}𝚙𝚔_2`$ iff, for all local name assignments $`l`$ consistent with $`w`$, we have $`w,l,𝚔_1_\mathrm{o}𝚙𝚔_2`$. The proof of this result (which, like that of all the technical results in this paper, is deferred to the appendix) uses standard techniques from the theory of fixed points. We now define our second semantics, called the closed semantics. It attempts to capture the intuition that the only bindings in $`l`$ should be those required by the certificates in $`c`$, using the minimal assignment promised by Theorem 3.1. We write $`w,𝚔_\mathrm{c}\varphi `$ if $`w,l_w,𝚔\varphi `$. We say that $`\varphi `$ is c-satisfiable if there exists a world $`w`$ and key $`𝚔`$ such that $`w,𝚔_\mathrm{c}\varphi `$ and that $`\varphi `$ is c-valid, denoted $`_\mathrm{c}\varphi `$, if $`w,𝚔_\mathrm{c}\varphi `$ for all worlds $`w`$ and principals $`𝚔`$. Note that by Theorem 3.1, the assignment $`l_w`$ is consistent with $`w`$, so c-satisfiability implies o-satisfiability. Thus, if $`_\mathrm{o}\varphi `$ then $`_\mathrm{c}\varphi `$. As we shall soon see (Theorem 3.5), somewhat surprisingly, the converse holds as well. ### 3.3 A Complete Axiomatization We start this section by presenting a sound and complete axiomatization for $`LLNC`$ with respect to the open semantics. We then prove that the open and closed semantics are characterized by the same valid formulas, so that the axiomatization is also sound and complete with respect to the closed semantics. The axiomatization depends in part on whether the set $`K`$ of keys is finite or infinite. Figure 3 describes the axiom system $`\text{AX}_{\mathrm{𝑖𝑛𝑓}}`$ for the case where $`K`$ is infinite. It is interesting to compare the axioms in $`\text{AX}_{\mathrm{𝑖𝑛𝑓}}`$ to Abadi’s axioms. Although we interpret $``$ as superset and he interprets it as subset, Reflexivity, Transitivity, Left-Monotonicity, and Associativity, hold in both cases, for essentially the same reasons. The switch from subset to superset means that the Converse of Globality holds in our case. Globality does not hold in general because the denotation of $`𝚙\text{’s}𝚐`$ may be empty if the denotation of $`𝚙`$ is empty (as we observed, this is also why the Converse of Globality does not hold in general for Abadi). In fact, for our logic, $`𝚙\text{’s}𝚐𝚐`$ holds whenever the interpretation of $`𝚙`$ is nonempty. We use $`𝚙\text{’s}𝚔𝚔`$ as a canonical way of denoting that the interpretation of $`𝚙`$ is nonempty. This explains the form of the Globality axiom. Since the interpretation of a key is always nonempty, we also get Key Globality. Key Linking is our analogue of Abadi’s Linking axiom. Of course, we use $`\mathrm{𝑐𝑒𝑟𝑡}`$ whereas Abadi uses $`\mathrm{𝑠𝑎𝑦𝑠}`$; in addition, only keys can certify formulas for us. While this axiom shows that there are some similarities between $`\mathrm{𝑐𝑒𝑟𝑡}`$ and $`\mathrm{𝑠𝑎𝑦𝑠}`$, there are some significant differences. We have no analogue of Abadi’s Speaking-for axiom and, unlike $`\mathrm{𝑠𝑎𝑦𝑠}`$, $`\mathrm{𝑐𝑒𝑟𝑡}`$ does not satisfy the standard axiom and rule of modal logic: $`(𝚔\mathrm{𝑐𝑒𝑟𝑡}(\varphi \psi ))(𝚔\mathrm{𝑐𝑒𝑟𝑡}\varphi )`$ does not imply $`𝚔\mathrm{𝑐𝑒𝑟𝑡}\psi `$ and $`𝚔\mathrm{𝑐𝑒𝑟𝑡}\varphi `$ is not valid even if $`\varphi `$ is valid. Interestingly, Abadi does not use these properties of Speaking-for in proving that his name resolution rules NR, used to capture REF2, are sound. As a result, (with very minor changes) we can show that the name resolution rules are also sound for LLNC, and hence we can prove analogues of Propositions 2.1 and 2.2. However, we can actually prove a much stronger result: whereas Abadi’s logic is able to draw conclusions about bindings that do not follow from REF2, LLNC captures REF2 exactly (see Theorem 4.1). $`\text{AX}_{\mathrm{𝑖𝑛𝑓}}`$ has two axioms that do not appear in Abadi’s axiomatization: Key Distinctness and Nonemptiness. Key Distinctness just captures the fact that we interpret keys as themselves. The first three parts of Nonemptiness capture various ways that an expression can be seen to be nonempty. For example, part (a) says that if $`𝚙`$ is bound to (i.e., is a superset of) a key, then its interpretation must be nonempty and part (b) says that if $`𝚙`$ is not a superset of $`𝚚`$, then $`𝚚`$ must be nonempty. Part (d) of Nonemptiness says that if $`𝚙`$ is nonempty and $`𝚔^{}`$ is bound to $`𝚙`$, then $`𝚙`$ is bound to $`𝚔^{}`$, i.e., $`𝚙`$ and $`𝚔^{}`$ have exactly the same interpretation. If $`K`$ is finite we need to add two further axioms to $`\text{AX}_{\mathrm{𝑖𝑛𝑓}}`$. Let $`\text{AX}_{\mathrm{𝑓𝑖𝑛}}`$ consist of all the axioms and rule in $`\text{AX}_{\mathrm{𝑖𝑛𝑓}}`$ together with: $$\begin{array}{cc}\text{Witnesses:}\hfill & \neg (𝚙𝚚)_{𝚔K}(\neg (𝚙𝚔)(𝚚𝚔))\hfill \\ & (𝚙\text{’s}𝚚)𝚔_1_{𝚔K}((𝚙𝚔)(𝚔\text{’s}𝚚𝚔_1))\hfill \\ \text{Current Principal:}\hfill & _{𝚔K}(𝚗_𝚔𝚕_𝚔𝚔\text{’s}𝚗_𝚔𝚕_𝚔)\hfill \\ & \text{where }𝚗_𝚔N\text{ and }𝚕_𝚔K\text{ for each }𝚔K\text{.}\hfill \end{array}$$ The two axioms that make up Witnesses essentially capture our interpretation of $``$ as containment. They tell us that facts about containment of principal expressions can be reduced to facts about keys. For example, the first one says that if $`𝚙`$ does not contain $`𝚚`$, then there is a key bound to $`𝚚`$ that is not bound to $`𝚙`$. Current Principal captures the fact that some key in $`K`$ must be the current principal; if $`𝚔`$ is the current principal, then for all local names $`𝚗`$ and keys $`𝚔^{}`$, $`𝚗𝚔^{}𝚔\text{’s}𝚗𝚔^{}`$ holds. (This is actually true not just for local names, but for all principal expressions; it suffices to state the axiom just for local names.) While the properties captured by these two axioms continue to hold even if $`K`$ is infinite, they can no longer be expressed in the logic, since we cannot take a disjunction over all the elements in $`K`$. Interestingly, we can drop Nonemptiness and Globality as axioms in $`\text{AX}_{\mathrm{𝑓𝑖𝑛}}`$. These properties already follow from the other properties in the presence of Witnesses. As the following result shows, these axiom systems completely characterize validity in the logic with respect to the open semantics. ###### Theorem 3.2 : $`\text{AX}_{\mathrm{𝑖𝑛𝑓}}`$ (resp., $`\text{AX}_{\mathrm{𝑓𝑖𝑛}}`$) is a sound and complete axiomatization of LLNC with respect to the open semantics if $`K`$ is infinite (resp., $`K`$ is finite). In the course of proving Theorem 3.2, we also prove a “finite model” result, which we cull out here. Let $`|\varphi |`$, the length of $`\varphi `$, be the total number of symbols appearing in $`\varphi `$. This result holds both when $`K`$ is finite and when $`K`$ is infinite. ###### Proposition 3.3 : Let $`K_\varphi `$ be the keys that appear in $`\varphi `$ and let $`C_\varphi (𝚔)`$ consist of all bindings $`𝚗𝚙`$ such that $`𝚔\mathrm{𝑐𝑒𝑟𝑡}𝚗𝚙`$ is a subformula of $`\varphi `$. If $`\varphi `$ is satisfiable with respect to the open semantics, then for all sets $`K^{}`$ of keys such that $`K_\varphi K^{}`$ and $`|K^{}|\mathrm{min}(|K|,2|\varphi |^2)`$, there is a world $`w=(\beta ,c)`$, local name assignment $`l`$, and principal $`𝚔K^{}`$ such that $`w,l,𝚔_\mathrm{o}\varphi `$ and (a) $`l(𝚔^{},𝚗)K^{}`$ for all $`𝚔^{}K`$ and $`𝚗N`$, (b) $`l(𝚔^{},𝚗)=\mathrm{}`$ if $`𝚔^{}K^{}`$, (c) $`\beta (𝚐)K^{}`$ for all $`𝚐G`$, (d) $`\beta (𝚐)=\mathrm{}`$ if $`𝚐`$ does not occur in $`\varphi `$, and (e) $`c(𝚔)C_\varphi (𝚔)`$ for all keys $`𝚔`$. ###### Corollary 3.4 : The problem of deciding if a formula $`\varphi LLNC`$ is satisfiable with respect to the open semantics is NP-complete (whether $`K`$ is finite or infinite). Proof: The lower bound is immediate from the fact that we can trivially embed satisfiability for propositional logic into satisfiability for LLNC. For the upper bound, given $`\varphi `$, choose $`K^{}`$ such that $`|K^{}|=\mathrm{min}(|K|,2|\varphi |^2)`$ and $`K^{}K_\varphi `$. Then guess $`w,l,𝚔`$ as in Proposition 3.3 and check whether $`w,l,𝚔_\mathrm{o}\varphi `$. Proposition 3.3 says that the guess is only polynomial in $`|\varphi |`$; it is clear that checking whether $`w,l,𝚔_\mathrm{o}\varphi `$ can also be done in time polynomial in $`\varphi `$. Note that for $`|\varphi ||K|`$ (which is likely to include all cases of practical interest, given that $`K`$ will typically be a very large set), the polynomial does not depend on $`|K|`$. As we suggested earlier, the closed semantics and the open semantics are characterized by exactly the same axioms. ###### Theorem 3.5 : The same formulas are c-valid and o-valid; i.e., for all formulas $`\varphi `$, we have $`_\mathrm{o}\varphi `$ iff $`_\mathrm{c}\varphi `$. We remark that this result is sensitive to the language under consideration. It may no longer hold if we move to a more expressive language. ###### Corollary 3.6 : $`\text{AX}_{\mathrm{𝑖𝑛𝑓}}`$ (resp., $`\text{AX}_{\mathrm{𝑓𝑖𝑛}}`$) is a sound and complete axiomatization of LLNC with respect to the closed semantics when $`K`$ is infinite (rep., finite). ###### Corollary 3.7 : The problem of deciding if a formula $`\varphi LLNC`$ is satisfiable with respect to the closed semantics is NP-complete (whether $`K`$ is finite or infinite). Let us now return to the contentious axioms discussed by Abadi. Converse of Globality is valid in $`LLNC`$, as we observed earlier. The generalization of Linking considered by Abadi, restricted to be syntactically well formed, amounts to $$(𝚔\mathrm{𝑐𝑒𝑟𝑡}(𝚙_1𝚙_2))(𝚔\text{’s}𝚙_1𝚔\text{’s}𝚙_2).$$ In general, this is not valid, since our semantics ignores certificates stating $`𝚙_1𝚙_2`$ when $`𝚙_1`$ is not a local name. Thus, we avoid the “unreasonable” conclusions that can be drawn from these axioms. In particular, it does not follow in our logic that $`(𝚔\mathrm{𝑐𝑒𝑟𝑡}(\mathrm{𝙳𝙽𝚂}!!𝚔))\mathrm{𝙳𝙽𝚂}!!𝚔`$. However, the reason it does not follow in LLNC is quite different from the reason it does not follow in Abadi’s logic: since $`\mathrm{𝙳𝙽𝚂}!!`$ is a global name, a certificate such as $`𝚔\mathrm{𝑐𝑒𝑟𝑡}(\mathrm{𝙳𝙽𝚂}!!𝚔)`$ has no impact on the interpretation of global names. This captures the intuition that $`𝚔`$ should not be trusted when making assertions about bindings not under its control. If we were willing to trust $`𝚔`$ on everything, then concluding that $`𝚔`$ is bound to $`\mathrm{𝙳𝙽𝚂}!!`$ after $`𝚔`$ certifies that it is would not seem so unreasonable. The following formula is also not valid in LLNC: $$(\neg (𝚔\mathrm{𝑐𝑒𝑟𝑡}\text{false})(𝚔\mathrm{𝑐𝑒𝑟𝑡}(\mathrm{𝙳𝙽𝚂}!!𝚔)))\mathrm{𝙳𝙽𝚂}!!𝚔.$$ (This formula corresponds to the one that we noted earlier is valid in Abadi’s logic.) Failure to issue a certificate stating false has no more impact on global names than does any other behavior of $`𝚔`$. Nor would a precondition asserting that the interpretation of $`𝚔`$ is non-empty validate the formula, since this is true in every world. We can in fact prove the following generalization of Abadi’s Proposition 2.5, which provides a stronger statement of the safety of our logic than Abadi’s result. ###### Proposition 3.8 : Let $`\mathrm{\Gamma }`$ be any c-satisfiable boolean combination of formulas of the form $`𝚔\mathrm{𝑐𝑒𝑟𝑡}\varphi `$, and let $`\mathrm{\Delta }`$ be any boolean combination of formulas of the form $`𝚙𝚚`$ where neither $`𝚙`$ nor $`𝚚`$ contains a local name. Then $`_\mathrm{c}\mathrm{\Gamma }\mathrm{\Delta }`$ iff $`_\mathrm{c}\mathrm{\Delta }`$. Informally, Proposition 3.8 says that facts about global names are completely independent of facts about certificates; issuing certificates can have no impact on the global name assignment. As we observed earlier, the analogous result does not hold for Abadi’s logic. ## 4 Name Resolution in LLNC In this section, we show that LLNC captures REF2 exactly. Indeed, we show that it does so for several distinct semantic interpretations. Define the order $``$ on worlds by $`(\beta ^{},c^{})(\beta ,c)`$ if 1. $`\beta ^{}(𝚐)\beta (𝚐)`$ for all global names $`𝚐`$, and 2. $`c^{}(𝚔)c(𝚔)`$ for all keys $`𝚔`$. That is, $`w^{}w`$ when $`w^{}`$ contains more certificates than $`w`$ and the bindings to global names in $`w`$ are a subset of those in $`w^{}`$. If $`E`$ is a set of formulas and $`\varphi `$ is a formula, we write $`E_\mathrm{o}\varphi `$ if for all worlds $`w`$, local name assigments $`l`$ consistent with $`w`$ and all keys $`𝚔`$, if $`w,l,𝚔_\mathrm{o}\psi `$ for all $`\psi `$ in $`E`$ then $`w,l,𝚔_\mathrm{o}\varphi `$. Similarly, $`E_\mathrm{o}\varphi `$ if for all worlds $`w`$ and all keys $`𝚔`$, if $`w,𝚔_\mathrm{c}\psi `$ for all $`\psi `$ in $`E`$ then $`w,𝚔_\mathrm{c}\varphi `$. ###### Theorem 4.1 : Suppose $`𝚔_1,𝚔_2`$ are principals, $`w=(\beta ,c)`$ is a world, and $`𝚙`$ is a principal expression. Let $`E_w`$ be the set of all formulas $`𝚐𝚔`$ for all global names $`𝚐`$ and keys $`𝚔\beta (𝚐)`$ and the formulas $`𝚔\mathrm{𝑐𝑒𝑟𝑡}\varphi `$ for all keys $`𝚔`$ and formulas $`\varphi c(𝚔)`$. The following are equivalent: 1. $`𝚔_1\text{REF2}(𝚔_2,\beta ,c,𝚙)`$, 2. $`w,𝚔_2_\mathrm{c}𝚙𝚔_1`$, 3. $`w^{},𝚔_2_\mathrm{c}𝚙𝚔_1`$ for all worlds $`w^{}w`$, 4. $`E_w_c𝚔_2\text{’s}𝚙𝚔_1`$, 5. $`E_w_o𝚔_2\text{’s}𝚙𝚔_1`$. This theorem gives a number of perspectives on name resolution in LLNC. The equivalence between (1) and (2) in this theorem tells us that REF2 is sound and complete with respect to key binding, according to the semantics of LLNC. That is, $`\text{REF2}(𝚔,\beta ,c,𝚙)`$ yields $`𝚔^{}`$ iff $`𝚙𝚔^{}`$ is forced to be true by the bindings of global names in $`\beta `$ and the certificates in $`c`$. Thus, viewed as a specification of the meaning of SDSI names, the closed semantics and REF2 are equivalent. Informally, we have viewed REF2 as a procedure that is run by an omniscient agent with complete information about the interpretation of global names and the certificates that have been issued. It is also possible to understand REF2 as performing a computation based on the limited information available to a particular principal. Suppose that the world $`w`$ expresses the limited information this principal has about the binding of global names and the certificates that have been issued. Suppose that $`w^{}`$ describes the actual bindings of global names and the certificates that have been issued. Assuming that all of the principal’s information is correct, then $`ww^{}`$. Thus, the set of $`w^{}w`$ is the set of all worlds $`w^{}`$ that are consistent with the information available to the principal. (We could formalize this using the Kripke semantics for the logic of knowledge in a distributed system \[HM90\].) The equivalence between (2) and (3) essentially shows that it doesn’t matter whether we view the principal as having total or partial information. The implication from (1) to (4) in Theorem 4.1 is analogous to Abadi’s soundness result, Proposition 2.2. Of course, the converse implication gives us completeness, which, as Abadi himself observed, does not hold for Abadi’s logic (since it validates conclusions that do not follow from REF2). Interestingly, although, as we have seen, there are significant differences between $`LLNC`$ and Abadi’s logic, an examination of Abadi’s soundness proof reveals that it does not use the Speaking-for rule, the unrestricted form of Globality, or the standard axiom and rule for the modal operator $`\mathrm{𝑠𝑎𝑦𝑠}`$, which are the main points of difference with our logic. This observation says that the proof of the implication from (1) to (4) is essentially the same for $`LLNC`$ and for Abadi’s logic. It is instructive to understand why the formulas considered in Examples 2.3 and 2.4, which give conclusions in Abadi’s logic beyond those derivable by REF2, are not valid in LLNC. It is easy to see why the formula $`𝚔\text{’s}(\mathrm{𝙻𝚊𝚖𝚙𝚜𝚘𝚗}\text{’s}𝚔^{})𝚔^{}`$ from Example 2.3 (which, by Associativity and Transitivity, is equivalent to $`(𝚔\text{’s}\mathrm{𝙻𝚊𝚖𝚙𝚜𝚘𝚗})\text{’s}𝚔^{}𝚔^{}`$) is not valid in LLNC. This is simply because the antecedent of (our version of) Globality does not always hold. Now consider the formula in Example 2.4. The proof that this is valid in Abadi’s logic uses the Speaking-for axiom, which does not hold for us (if we replace $`\mathrm{𝑠𝑎𝑦𝑠}`$ by $`\mathrm{𝑐𝑒𝑟𝑡}`$). To see that it is not valid in LLNC, consider a world $`w=(\beta ,c)`$ containing only the certificates forced by the formulas (i.e., $`c(𝚔)=\{\mathrm{𝙻𝚊𝚖𝚙𝚜𝚘𝚗}𝚔_1,\mathrm{𝙻𝚊𝚖𝚙𝚜𝚘𝚗}𝚔_2\}`$, $`c(𝚔_1)=\{\mathrm{𝚁𝚘𝚗}\mathrm{𝚁𝚒𝚟𝚎𝚜𝚝}\}`$, $`c(𝚔_2)=\{\mathrm{𝚁𝚒𝚟𝚎𝚜𝚝}𝚔_3\}`$). Then it is easy to see that $`w,𝚔\vDash ̸𝚔\text{’s}(\mathrm{𝙻𝚊𝚖𝚙𝚜𝚘𝚗}\text{’s}\mathrm{𝚁𝚘𝚗})𝚔_3`$, since $`[[𝚔\text{’s}(\mathrm{𝙻𝚊𝚖𝚙𝚜𝚘𝚗}\text{’s}\mathrm{𝚁𝚘𝚗})]]_{w,l_w,𝚔}=\mathrm{}`$ whereas $`[[𝚔_3]]_{w,l_w,𝚔}=\{𝚔_3\}`$. ## 5 Logic Programming Implementations of Name Resolution Queries The reader familiar with the theory of logic programming may have noted a close resemblance of the results and constructions of the preceding sections to the (now standard) fixpoint semantics for logic programs developed originally by van Emden and Kowalski \[EK76\]. Indeed, it is possible to translate our semantics into the framework of logic programming. In fact, we provide a translation that does not require the use of function symbols and thus produces a Datalog program, a restricted type of logic program that has significant computational advantages over unrestricted logic programs. Our translation allows us to take advantage of the significant body of research on the optimization of Datalog programs \[Ull88, Ull89\]. The idea is to translate queries to formulas in a first-order language over a vocabulary $`V`$ which consists of a constant symbol for each element in $`KGN`$ and a ternary predicate symbol $`\mathrm{𝚗𝚊𝚖𝚎}`$. Intuitively, $`\mathrm{𝚗𝚊𝚖𝚎}(x,y,z)`$ says that, in the local name space of key $`x`$, the basic principal expression (i.e., key, global name or local name) $`y`$ is bound to key $`z`$. Using $`\mathrm{𝚗𝚊𝚖𝚎}`$, for each principal expression $`𝚙`$ and pair of variables $`x,y`$, we define a first-order formula $`\tau _{x,y}(𝚙)`$ that, intuitively, corresponds to the assertion “$`y[[𝚙]]_x`$,” by induction on the structure of $`𝚙`$: 1. $`\tau _{x,y}(𝚙)=\mathrm{𝚗𝚊𝚖𝚎}(x,𝚙,y)`$ when $`𝚙KGN`$. 2. $`\tau _{x,y}(𝚚\text{’s}𝚛)=z(\tau _{x,z}(𝚚)\tau _{z,y}(𝚛))`$, where $`zx,y`$. Recall that a Herbrand structure over the vocabulary $`V`$ is a first-order structure that has as its domain the set of constant symbols $`KGN`$ in $`V`$ and interprets each constant sybol as itself. Such a structure may be represented as a set of tuples of the form $`\mathrm{𝚗𝚊𝚖𝚎}(x,y,z)`$, where $`x,y,zKGN`$. The subset relation on such sets partially orders the Herbrand structures. We say that a Herbrand structure $`M`$ over $`V`$ represents a world $`w=(\beta ,c)`$ and local name assignment $`l`$ if, for all $`x,y,zKGN`$, we have $`\mathrm{𝚗𝚊𝚖𝚎}(x,y,z)M`$ iff either 1. $`x,y,zK`$ and $`z=y`$, or 2. $`xK`$, $`yG`$ and $`z\beta (y)`$, or 3. $`xK`$, $`yN`$ and $`zl(x,y)`$. Intuitively, $`M`$ represents $`w`$ and $`l`$ if it encodes all the interpretations of basic principal expressions given by $`w`$ and $`l`$. The following result, whose straightforward proof is left to the reader, shows that in this case $`M`$ also captures the interpretation of all other principal expressions, and expresses the correctness of our translation of principal expressions. ###### Proposition 5.1 : If $`M`$ represents $`w`$ and $`l`$ then, for all principal expressions $`𝚙`$ and $`x,yKGN`$, we have $`M\tau _{x,y}(𝚙)`$ iff $`x,yK`$ and $`w,l,x𝚙y`$. We now show how a logic program can be used to capture the relationsip between $`w`$ and $`l_w`$. For each world $`w=(\beta ,c)`$, we define a theory (set of sentences) $`\mathrm{\Sigma }_w`$ that characterizes $`w`$; $`\mathrm{\Sigma }_w`$ consists of the following sentences: 1. a sentence $`\mathrm{𝚗𝚊𝚖𝚎}(𝚔_1,𝚔_2,𝚔_2)`$, for each pair of keys $`𝚔_1,𝚔_2K`$, and 2. the sentence $`\mathrm{𝚗𝚊𝚖𝚎}(𝚔_1,𝚐,𝚔_2)`$, for each pair of keys $`𝚔_1,𝚔_2K`$ and global name $`𝚐G`$ such that $`𝚔_2\beta (𝚐)`$, 3. the sentence $`y(\tau _{𝚔,y}(𝚚)\mathrm{𝚗𝚊𝚖𝚎}(𝚔,𝚗,y))`$, for each key $`𝚔`$ and binding $`𝚗𝚚`$ in $`c(𝚔)`$. After some equivalence-preserving syntactic transformations (moving the existentials in the body of these sentences to the front), the theory $`\mathrm{\Sigma }_w`$ is a definite Horn theory, i.e., it consists of formulas of the form $`𝐱(BH)`$, where $`B`$ is a (possibly empty) conjunction of atoms (that is, formulas of the form $`\mathrm{𝚗𝚊𝚖𝚎}(x,y,z)`$ or $`y=z`$) and $`H`$ is an atom. Well-known results from the theory of logic programming show that such a theory $`\mathrm{\Sigma }`$ has a Herbrand model $`M_\mathrm{\Sigma }`$ minimal with respect to the containment ordering on Herbrand structures. Moreover, this minimal Herbrand model captures the minimal name assignments for $`w`$. ###### Theorem 5.2 : The minimal Herbrand model $`M_w`$ of $`\mathrm{\Sigma }_w`$ represents $`w`$ and $`l_w`$. Using Proposition 5.1, we immediately obtain the following corollary. ###### Corollary 5.3 : For all $`x,yKGN`$ and principal expressions $`𝚙`$, we have $`M_w\tau _{x,y}(𝚙)`$ iff $`x,yK`$ and $`w,x_\mathrm{c}𝚙y`$. Because $`\mathrm{\Sigma }_w`$ is a definite Horn theory, it corresponds to a logic program. Moreover, for existential queries, i.e., queries $`\varphi `$ that are sentences formed from atomic formulas using only conjunction, disjunction and existential quantification (but not negation), we have that $`\mathrm{\Sigma }`$ entails $`\varphi `$ iff $`M_\mathrm{\Sigma }\varphi `$. This enables us to exploit logic programming technology to obtain efficient implementations of several types of queries, corresponding to different choices of bound and free variables in the predicate “$`\mathrm{𝚗𝚊𝚖𝚎}`$”. We may even form complex queries not corresponding in any direct way to the capacities of the procedure REF2. Examples of this include the following: 1. the query $`\mathrm{𝚗𝚊𝚖𝚎}(𝚔_1,𝚗,𝚔_2)`$ returns “yes” if $`𝚔_2`$ is bound to the local name $`𝚗`$ according to $`𝚔_1`$; 2. the query $`\mathrm{𝚗𝚊𝚖𝚎}(X,𝚗,𝚔)`$ returns the set of keys $`X`$ such that $`𝚔`$ is in $`𝚗`$ according to $`X`$; 3. the query $`\mathrm{𝚗𝚊𝚖𝚎}(𝚔_1,X,𝚔_2)`$ returns the set of global and local names $`X`$ containing $`𝚔_2`$ according to $`𝚔_1`$. 4. the query $`\mathrm{𝚗𝚊𝚖𝚎}(𝚔_1,𝚗,X)\mathrm{𝚗𝚊𝚖𝚎}(𝚔_2,𝚗,X)`$ returns the set of keys $`X`$ that $`𝚔_1`$ and $`𝚔_2`$ agree to be associated with local name $`𝚗`$. Many more possibilities clearly exist. These observations show the advantage of viewing name resolution in a logic programming framework. ## 6 Self Abadi considers an extension of his logic obtained by adding a special basic principal expression $`\mathrm{𝚂𝚎𝚕𝚏}`$, intended to represent SDSI’s expression (ref:). (We remark that $`\mathrm{𝚂𝚎𝚕𝚏}`$ is essentially the same as $`I`$ in the logic of naming considered in \[GH93\].) Intuitively, $`\mathrm{𝚂𝚎𝚕𝚏}`$ denotes the current principal. The semantics given to $`\mathrm{𝚂𝚎𝚕𝚏}`$ by Abadi extends the definition of the set of principals associated with a principal expression by taking $`[[\mathrm{𝚂𝚎𝚕𝚏}]]_a=\{a\}`$ for each $`a𝒲`$. This suffices to validate the following axiom. $$\begin{array}{ccc}\text{Identity:}\hfill & \mathrm{𝚂𝚎𝚕𝚏}\text{’s}𝚙𝚙\hfill & 𝚙\mathrm{𝚂𝚎𝚕𝚏}\text{’s}𝚙\hfill \\ & 𝚙\text{’s}\mathrm{𝚂𝚎𝚕𝚏}𝚙\hfill & 𝚙𝚙\text{’s}\mathrm{𝚂𝚎𝚕𝚏}\hfill \end{array}$$ These axioms very reasonably capture the intuitions that $`\mathrm{𝚂𝚎𝚕𝚏}`$ refers to the current principal. However, not all consequences of this semantics for $`\mathrm{𝚂𝚎𝚕𝚏}`$ are so reasonable. For example, the following is valid under Abadi’s semantics: $$\begin{array}{c}(𝚔_P\mathrm{𝑠𝑎𝑦𝑠}\mathrm{𝚄𝚂}\mathrm{𝚂𝚎𝚕𝚏})(𝚔_P\mathrm{𝑠𝑎𝑦𝑠}\mathrm{𝚄𝚂}𝚔_{VP})\hfill \\ 𝚔_P\mathrm{𝑠𝑎𝑦𝑠}((\mathrm{𝚄𝚂}\mathrm{𝑠𝑎𝑦𝑠}\mathrm{𝑓𝑎𝑙𝑠𝑒})(\mathrm{𝚂𝚎𝚕𝚏}𝚔_{VP}))\hfill \end{array}$$ (1) Interpreting $`𝚔_P`$ as the key of the president of the US and $`𝚔_{VP}`$ the key of the vice-president, this is clearly unreasonable. It should not follow from the fact the the president says that both he and the vice-president speak for the US that according to the president, either the US speaks nonsense or the vice president speaks for the president. Abadi’s suggested semantics for $`\mathrm{𝚂𝚎𝚕𝚏}`$ works much better in the context of the logic LLNC. Suppose we extend this logic to include $`\mathrm{𝚂𝚎𝚕𝚏}`$, and like Abadi, define $`[[\mathrm{𝚂𝚎𝚕𝚏}]]_𝚔=\{𝚔\}`$ for keys $`𝚔K`$. This again validates the Identity axioms above. To get completeness, we just need to add one axiom in addition to Identity, which basically says that $`\mathrm{𝚂𝚎𝚕𝚏}`$ acts like a key (cf. Nonemptiness (d)): $$\text{Self-is-key}\mathrm{𝚂𝚎𝚕𝚏}𝚙𝚙\text{’s}𝚔𝚔𝚙\mathrm{𝚂𝚎𝚕𝚏}.$$ Let $`\text{AX}_{\mathrm{𝑖𝑛𝑓}}^{\mathrm{𝑠𝑒𝑙𝑓}}`$ (resp., $`\text{AX}_{\mathrm{𝑓𝑖𝑛}}^{\mathrm{𝑠𝑒𝑙𝑓}}`$) be the result of adding Identity and Self-is-key to $`\text{AX}_{\mathrm{𝑖𝑛𝑓}}`$ (resp., $`\text{AX}_{\mathrm{𝑓𝑖𝑛}}`$). Let LLNC<sup>s</sup> be the language that results when we add $`\mathrm{𝚂𝚎𝚕𝚏}`$ to the syntax. ###### Theorem 6.1 : $`\text{AX}_{\mathrm{𝑖𝑛𝑓}}^{\mathrm{𝑠𝑒𝑙𝑓}}`$ (resp., $`\text{AX}_{\mathrm{𝑓𝑖𝑛}}^{\mathrm{𝑠𝑒𝑙𝑓}}`$) is a sound and complete axiomatization of LLNC<sup>s</sup> with respect to the open semantics if $`K`$ is infinite (resp., $`K`$ is finite). Propositions 3.3 and Theorem 3.5 hold with essentially no change in proof for LLNC<sup>s</sup>; it follows that $`\text{AX}_{\mathrm{𝑓𝑖𝑛}}^{\mathrm{𝑠𝑒𝑙𝑓}}`$ (resp., $`\text{AX}_{\mathrm{𝑖𝑛𝑓}}^{\mathrm{𝑠𝑒𝑙𝑓}}`$) is also complete with respect to the closed semantics and the satisfiability problem is NP-complete. Interestingly, the proof of completeness shows that once we add Identity and Self-is-key to the axioms, we no longer need Current Principal as an axiom in the finite case. Here is a sketch of the argument: From Identity we get that $`\mathrm{𝚂𝚎𝚕𝚏}\text{’s}𝚔𝚔`$ is provable for any key $`𝚔`$. Now applying Witnesses, we get that $`_{𝚔K}\mathrm{𝚂𝚎𝚕𝚏}𝚔`$ is provable. Together with Self-is-key, this says that $`\mathrm{𝚂𝚎𝚕𝚏}`$ is one of the keys in $`K`$. Identity (together with Transitivity) tells us that for that key $`𝚔`$ that is Self, $`𝚗𝚔^{}𝚔\text{’s}𝚗𝚔^{}`$ holds, giving us Current Principal. Note that with our semantics for $`\mathrm{𝚂𝚎𝚕𝚏}`$, the counterintuitive conclusion (1) does not follow. From $`𝚔_P\mathrm{𝑐𝑒𝑟𝑡}\mathrm{𝚄𝚂}\mathrm{𝚂𝚎𝚕𝚏}`$ and $`𝚔_P\mathrm{𝑐𝑒𝑟𝑡}\mathrm{𝚄𝚂}𝚔_{VP}`$ it follows that $`[[\mathrm{𝚄𝚂}]]_{𝚔_P}\{𝚔_P,𝚔_{VP}\}`$. Thus, we have neither $`[[\mathrm{𝚄𝚂}]]_{𝚔_P}=\mathrm{}`$ nor $`\{𝚔_{VP}\}[[\mathrm{𝚄𝚂}]]_{𝚔_P}`$, which would be required to get a conclusion similar to that drawn by Abadi’s logic. ## 7 Conclusions We have introduced a logic LLNC for reasoning about SDSI’s local name spaces and have argued that it has some significant advantages over Abadi’s logic. Among other things, it provides a complete characterization of SDSI’s REF2, has an elegant complete axiomatization, and its connections with Logic Programming lead to efficient implementations of many queries of interest. We believe that some of the dimensions in which Abadi’s logic differs from SDSI warrant further investigation. For example, under some sensible interpretations, the conclusions reached by Abadi’s logic in Example 2.4 are quite reasonable. One such interpretation is that while local names may be bound to more than one key, they are intended to denote a single individual. If $`𝚔`$ knows that $`𝚔_1`$ and $`𝚔_2`$ are two keys used by the one individual Lampson, and Lampson uses $`𝚔_1`$ to certify that his local name Ron is bound to the name Rivest, and also uses his key $`𝚔_2`$ to certify that his local name Rivest is bound to $`𝚔_3`$, then it is very reasonable to conclude that $`𝚔\text{’s}\mathrm{𝙻𝚊𝚖𝚙𝚜𝚘𝚗}\text{’s}\mathrm{𝚁𝚘𝚗}`$ is bound to $`𝚔_3`$. Another interpretation supporting this conclusion would be that $`\mathrm{𝑠𝑎𝑦𝑠}`$ aggregates the certificates issued using a number of distinct keys (possibly belonging to distinct individuals) much in the way that the notion of distributed knowledge \[FHMV95\] from the literature on reasoning about knowledge aggregates the knowledge of a collection of agents. We believe that our semantic framework, which, unlike Abadi’s, makes the set of certificates issued explicit, provides an appropriate basis for the study of such issues. Our semantic framework also lends itself to a number of generalizations, which we are currently exploring. These include reasoning about the beliefs of principals and reasoning about permission, authority, and delegation. We hope to report on this work shortly. ## Appendix A Proofs In this appendix, we prove all the technical results stated in the main text. For ease of exposition, we repeat the statements of the results here. Theorem 3.1: Given a world $`w`$, there exists a unique local name assignment $`l_w`$ minimal in the set of all local name assignments consistent with $`w`$. Moreover, if $`𝚙`$ is a principal expression and $`𝚔_1`$ and $`𝚔_2`$ are keys, then $`w,l_w,𝚔_1_\mathrm{o}𝚙𝚔_2`$ iff, for all local name assignments $`l`$ consistent with $`w`$, we have $`w,l,𝚔_1_\mathrm{o}𝚙𝚔_2`$. Proof: This result can be established using standard results from the theory of fixed points. Suppose $`(X,)`$ is a complete partial order. Denote the least upper bound of a set $`YX`$ by $`Y`$. A mapping $`T:XX`$ is said to be monotonic if for all $`xy`$ in $`X`$ we have $`T(x)T(y)`$. Such a mapping $`T`$ is said to be continuous if for all infinite increasing sequences $`x_0x_1\mathrm{}`$ in $`X`$ we have $`T(\{x_i:i𝐍\})=\{T(x_i):i𝐍\}`$. Note that continuity implies monotonicity. To establish continuity of a monotonic mapping $`T`$, it suffices to show that $`T(\{x_i:i𝐍\})\{T(x_i):i𝐍\}`$, since the opposite containment is immediate from monotonicity. For a fixed expression $`𝚙`$, world $`w`$ and key $`𝚔`$, the expression $`[[𝚙]]_{w,l,𝚔}`$ is easily seen to be monotonic in $`l`$, i.e., if $`ll^{}`$ then $`[[𝚙]]_{w,l,𝚔}[[𝚙]]_{w,l^{},𝚔}`$. Moreover, it is also continuous in $`l`$. ###### Lemma A.1 : Suppose $`l_0l_1\mathrm{}`$ is an increasing sequence of local name assignments and let $`l_\omega =_{m𝐍}l_m`$. For all principal expressions $`𝚙`$, we have $`[[𝚙]]_{w,l_\omega ,𝚔}=_{m𝐍}[[𝚙]]_{w,l_m,𝚔}`$. Proof: By a straightforward induction on the structure of $`𝚙`$. Given the world $`w=(\beta ,c)`$, we define an operator $`T_w`$ on the space of local name assignments $`\mathrm{𝐿𝑁𝐴}`$. For a local name assignment $`l`$, we define $`T_w(l)`$ to be the local name assignment such that for all $`𝚔K`$ and $`𝚗N`$, the set $`T_w(l)(𝚔,𝚗)`$ is the union of the sets $`[[𝚙]]_{w,l,𝚔}`$ such that the formula $`𝚗𝚙`$ is in $`c(𝚔)`$. The following lemma is follows easily from Lemma A.1. ###### Lemma A.2 : The mapping $`T_w`$ is a continuous operator on $`(\mathrm{𝐿𝑁𝐴},)`$. The following lemma is almost immediate from the definitions. ###### Lemma A.3 : A local name assignment $`l`$ is consistent with a world $`w`$ iff $`T_w(l)l`$. Suppose $`(X,)`$ is a complete partial order with minimal element $``$. An element $`xX`$ is said to be a pre-fixpoint of an operator $`T`$ on $`X`$ if $`T(x)x`$; $`x`$ is a fixpoint of $`T`$ if $`T(x)=x`$. Given an operator $`T`$ on $`X`$, define a sequence of elements $`T\gamma `$, where $`\gamma `$ is an ordinal, as follows. For the base case, let $`T0=`$. For successor ordinals $`\gamma +1`$, define $`T\gamma +1=T(T\gamma )`$. For limit ordinals $`\gamma `$, define $`T\gamma =\{T\delta :\delta <\gamma \}`$. A well-known result (see \[LNS82\] for a discussion of its history) states that if $`T`$ is continuous then then this sequences converges to the least pre-fixpoint of $`T`$, that convergence has taken place by $`\gamma =\omega `$, and that $`T\omega `$ is in fact a fixed point of $`T`$. Thus, we obtain as a corollary of Lemma A.2 and Lemma A.3 that there exists a minimal local name assignment consistent with $`w`$, and that this local name assignment equals $`T_w\omega `$. The second half of Theorem 3.1 is immediate from the earlier observation that $`[[𝚙]]_{w,l,𝚔}`$ is monotonic in $`l`$. Theorem 3.2: $`\text{AX}_{\mathrm{𝑓𝑖𝑛}}`$ (resp., $`\text{AX}_{\mathrm{𝑖𝑛𝑓}}`$) is a sound and complete axiomatization of LLNC with respect to the open semantics if $`K`$ is infinite (resp., $`K`$ is finite). Proof: We start with the completeness proof for $`\text{AX}_{\mathrm{𝑖𝑛𝑓}}`$, so that we assume that $`K`$ is infinite. We then show how to deal with $`\text{AX}_{\mathrm{𝑓𝑖𝑛}}`$. As usual, it suffices to show that if $`\varphi `$ is $`\text{AX}_{\mathrm{𝑖𝑛𝑓}}`$-consistent, then $`\varphi `$ is satisfable. In fact, we put a little extra work into our proof that $`\varphi `$ is satisfiable so that we can prove Proposition 3.3 as well. Let $`\text{Sub}(\varphi )`$ consist of all subformulas of $`\varphi `$. We say that a principal expression $`𝚙^{}`$ is a variant of $`𝚙`$ if $`𝚙𝚙^{}`$ and $`𝚙^{}𝚙`$ are both provable using only Reflexivity, Associativity, and Transitivity. The left-associative variant of a principal expression $`𝚙`$ is the one where we associate all terms to the left. Thus, $`((𝚗_1\text{’s}𝚗_2)\text{’s}𝚗_3)\text{’s}𝚗_4`$ is the left-associative variant of $`𝚗_1\text{’s}((𝚗_2\text{’s}𝚗_3)\text{’s}𝚗_4)`$. Define $`P`$ to be the smallest set of principal expressions such that 1. if $`𝚙𝚚`$ is in $`\text{Sub}(\varphi )`$ then $`𝚙`$ and $`𝚚`$ are in $`P`$, 2. if $`𝚔\mathrm{𝑐𝑒𝑟𝑡}(𝚗𝚙)\text{Sub}(\varphi )`$ then $`𝚔\text{’s}𝚗`$ and $`𝚔\text{’s}𝚙`$ are in $`P`$, 3. if $`𝚙P`$ and $`𝚙^{}`$ is the left-associative variant of $`𝚙`$, then $`𝚙^{}P`$, 4. $`P`$ is closed under subexpressions, so that if $`𝚙\text{’s}𝚚P`$, then so are $`𝚙`$ and $`𝚚`$, 5. if $`𝚔P`$ is a key and $`𝚗P`$ is a local name, then $`𝚔\text{’s}𝚗P`$. For Proposition 3.3, it is necessary to get an upper bound on the size of $`P`$ in terms of $`|\varphi |`$. ###### Lemma A.4 : $`|P|<2|\varphi |^2`$. Proof: Let $`|𝚙|`$ be the total number of expressions in $`GKN`$ that appear in $`𝚙`$, counted with multiplicity. An easy proof by induction on structure shows that a principal expression $`𝚙`$ has at most $`|𝚙|`$ subexpressions, at least one of which must be in $`GKN`$. For every other subexpression $`𝚚`$, there is a unique left-associative variant $`𝚚^{}`$, which has at most $`|𝚚^{}|=|𝚚||𝚙|`$ subexpressions, each of which is associated to the left. Thus, starting with a principal expression $`𝚙`$, the least set closed under clauses 3 and 4 above contains at most $`|𝚙|^2`$ elements. Now a straightforward induction on the structure of $`\varphi `$ shows that the least set $`P^{}`$ closed under clauses 1-4 above has at most $`|\varphi |^2`$ expressions. Finally, it is easy to see that closing off under 5 gives us $`P`$, since the set that results after closing off under 5 is still closed under 1–4. Moreover, this final step adds at most $`|\varphi |^2`$ expressions $`𝚔\text{’s}𝚗`$, since both $`𝚔`$ and $`𝚗`$ must be subexpressions of $`\varphi `$. Let $`𝚔_0`$ be some key not occurring in $`P`$. We use $`𝚔_0`$ both to express emptiness of expressions in $`P`$ and as the “current principal”. Define $`P_1`$ to be the set of principal expressions $`P\{𝚔_0\}\{𝚙\text{’s}𝚔_0:𝚙P\}`$. Let $`E`$ be consist of the formulas $`𝚙\text{’s}𝚔_0𝚔_0`$ for each $`𝚙P`$. Note that all principal expressions occurring in the formulas in $`E`$ are in $`P_1`$. Let $`S`$ be an $`\text{AX}_{\mathrm{𝑖𝑛𝑓}}`$-consistent set containing $`\varphi `$ and, for every formula $`\psi \text{Sub}(\varphi )E`$, either $`\psi `$ or $`\neg \psi `$. Since $`\varphi `$ is $`\text{AX}_{\mathrm{𝑖𝑛𝑓}}`$-consistent, there must be some $`\text{AX}_{\mathrm{𝑖𝑛𝑓}}`$-consistent set $`S`$ of this form. Define $`S^+=Cl(S,P_1)`$ to be the smallest set of formulas containing $`S`$ closed under Reflexivity, Transitivity, Left Motonocity, Converse of Globality, Globality, and Nonemptiness, in the sense that * if $`𝚙P_1`$, then $`𝚙𝚙S^+`$, * if $`𝚙𝚚`$ and $`𝚚𝚛`$ are both in $`S^+`$, then $`𝚙𝚛S^+`$, * if $`𝚙𝚚S^+`$, $`𝚙\text{’s}𝚛P_1`$, and $`𝚚\text{’s}𝚛P_1`$, then $`𝚙\text{’s}𝚛𝚚\text{’s}𝚛S^+`$, * if $`𝚙\text{’s}𝚐P_1`$ for $`𝚐KG`$ then $`𝚐𝚙\text{’s}𝚐S^+`$, * if $`𝚙\text{’s}𝚔𝚔S^+`$ for some key $`𝚔`$ and $`𝚙\text{’s}𝚐P_1`$, where $`𝚐KG`$, then $`𝚙\text{’s}𝚐𝚐S^+`$, * if $`𝚔\mathrm{𝑐𝑒𝑟𝑡}(𝚗𝚙)S^+`$ then $`(𝚔\text{’s}𝚗𝚔\text{’s}𝚙)S^+`$, * if $`𝚙𝚔^{}S^+`$ and $`𝚙\text{’s}𝚔P_1`$, then $`𝚙\text{’s}𝚔𝚔S^+`$, * if $`\neg (𝚙𝚚)S^+`$ and $`𝚚\text{’s}𝚔P_1`$, then $`𝚚\text{’s}𝚔𝚔S^+`$, * if $`𝚙\text{’s}𝚚𝚔_1S^+`$ and $`𝚙\text{’s}𝚔P_1`$, then $`𝚙\text{’s}𝚔𝚔S^+`$, * if $`𝚙\text{’s}𝚔𝚔`$ and $`𝚔^{}𝚙`$ are both in $`S^+`$, then $`𝚙𝚔^{}S^+`$, * if $`𝚔`$ and $`𝚔^{}`$ are distinct keys in $`P`$, then $`\neg (𝚔𝚔^{})S^+`$, * If $`𝚙^{}`$ is the left-associative variant of $`𝚙P`$, then $`𝚙𝚙^{}S^+`$ and $`𝚙^{}𝚙S^+`$. It is easy to see that $`S^+`$ is $`\text{AX}_{\mathrm{𝑖𝑛𝑓}}`$-consistent, since $`S`$ is and each of the closure rules emulates an axiom in $`\text{AX}_{\mathrm{𝑖𝑛𝑓}}`$. Our goal now is to show that there exists a triple $`w,l,𝚔`$ such that $`w,l,𝚔\psi `$ for all $`\psi S`$ (and thus, in particular, $`w,l,𝚔\varphi `$). ###### Lemma A.5 : If $`𝚔_0`$ appears in the formula $`𝚙𝚚S^+`$, then $`𝚔_0`$ appears in both $`𝚙`$ and $`𝚚`$. Proof: An easy induction on the construction of $`S^+`$, using the fact that all principal expressions occurring in $`S^+`$ are in $`P_1`$ and $`𝚔_0`$ appears only as the right most expression in a principal expression in $`P_1`$. By Lemma A.5, if $`𝚙𝚚S^+`$ and one of the expressions $`𝚙,𝚚`$ is in $`P`$ (and thus does not mention $`𝚔_0`$) then so is the other. Define a binary relation $``$ on $`P`$ by defining $`pq`$ if both $`𝚙𝚚`$ and $`𝚚𝚙`$ are in $`S^+`$. It is immediate from transitivity and reflexivity that $``$ is an equivalence relation on $`P`$. Given $`𝚙P`$, we write $`[𝚙]`$ for the equivalence class of $`𝚙`$ under $``$. We classify the expressions in $`P`$ as follows. Say that an expression $`𝚙`$ in $`P`$ is empty (with respect to $`S^+`$) if $`\neg (𝚙\text{’s}𝚔_0𝚔_0)`$ is in $`S^+`$. Say that $`𝚙`$ is key-equivalent if it is not empty and $`𝚔𝚙`$ is in $`S^+`$ for some key $`𝚔`$ (by (ClNE) this implies $`𝚙𝚔`$). Intuitively, the interpretation of an empty expression will be the empty set and the interpretation of a key-equivalent expression $`𝚙`$ such that $`𝚔𝚙S^+`$ will be $`\{𝚔\}`$. If $`𝚙`$ is neither empty nor key-equivalent, we say it is open. Clearly, every expression in $`P`$ is either empty, key-equivalent, or open. Moreover, by (ClLM) and (ClT), if $`𝚙𝚚`$ then $`𝚙`$ is empty, key-equivalent or open iff $`𝚚`$ is. In particular, we may sensibly refer to open $``$-equivalence classes of expressions in $`P`$. Let $`O`$ be the set of open equivalence classes of expressions in $`P`$. Note that if $`K_\varphi K`$ consists of all the keys in $`K`$ that appear in $`\varphi `$, then there are fewer than $`2|\varphi |^2|K_\varphi |`$ equivalence classes of open expressions. For each class $`cO`$, let $`𝚔_c`$ be a fresh key. Intuitively, the key $`𝚔_c`$ will act as a canonical representative of the keys in the interpretation of an expression $`𝚙c`$, in the sense that the interpretations of $`𝚙\text{’s}𝚚`$ and $`𝚔_c\text{’s}𝚚`$ will be the same for certain expressions $`𝚚`$. Since $`K`$ is infinite, we are guaranteed that we can always find keys $`𝚔_c`$, but the argument works even if $`K`$ is finite, as long as $`|K|2|\varphi |^2`$. (We also need to have a key in $`KK_\varphi `$ to be $`𝚔_0`$.) Define $`S^{}`$ to be consist of $`S^+`$ together with, for all $`cO`$, 1. the formula $`𝚔_c𝚔_c`$, and 2. the formulas $`𝚙𝚔_c`$, where for some $`𝚚c`$ we have $`𝚙𝚚S^+`$. It is easy to show that $`𝚔_0`$ does not appear in any formula in $`S^{}S^+`$: Clearly $`𝚔_0`$ does not appear in the formulas $`𝚔_c𝚔_c`$ added by clause 1. If $`𝚙𝚚`$ is a formula added by clause 2, then there is some equivalence class $`c`$ and expression $`𝚚c`$ such that $`𝚙𝚚S^+`$. Since $`c`$ is an equivalence class of expressions in $`P`$, none of which contain $`𝚔_0`$, the expression $`𝚚`$ does not contain $`𝚔_0`$. It follows from Lemma A.5 that $`𝚙`$ does not contain $`𝚔_0`$. Since $`S^{}S^+`$ contains no formulas involving $`𝚔_0`$, $`S^{}`$ also satisfies the property stated for $`S^+`$ in Lemma A.5. Define the local name assignment $`l`$ as follows. Given a key $`𝚔`$ and local name $`𝚗`$, 1. $`l(𝚔_0,𝚗)=\{𝚔^{}K|𝚗𝚔^{}S^{}\}`$, 2. $`l(𝚔,𝚗)=\{𝚔^{}K|𝚔\text{’s}𝚗𝚔^{}S^{}\}`$ if $`𝚔P`$, 3. $`l(𝚔,𝚗)=\{𝚔^{}K|𝚙\text{’s}𝚗𝚔^{}S^{}\text{and}𝚙c\}`$ if $`𝚔=𝚔_c`$ for some $`cO`$, 4. $`l(𝚔,𝚗)=\mathrm{}`$ for all other $`𝚔`$. Define the world $`w=(\beta ,c)`$ by taking $`\beta (𝚐)=\{𝚔K|𝚐𝚔S^{}\}`$ and defining $`c(𝚔)`$, for each key $`𝚔`$, to be the set of formulas $`𝚗𝚙`$ such that $`(𝚔\mathrm{𝑐𝑒𝑟𝑡}(𝚗𝚙))S`$. Note for future reference that there exists a finite subset $`K_1`$ of $`K`$ such that $`l(𝚗,𝚔)K_1`$, $`l(𝚗,𝚔)=\mathrm{}`$ for $`𝚔K_1`$, $`\beta (𝚐)K_1`$, and $`\beta (𝚐)=\mathrm{}`$ if $`𝚐`$ does not appear in $`\varphi `$. Indeed, $`K_1`$ consists of the keys that appear in $`S`$, $`𝚔_0`$, and the keys $`𝚔_c`$ for $`cO`$. Let $`I(𝚙)=\{𝚔K|𝚙𝚔S^{}\}`$. ###### Lemma A.6 : If $`𝚙P`$, then $`𝚙`$ is empty iff $`I(𝚙)=\mathrm{}`$. Proof: If $`𝚙`$ is not empty, then it is either key-equivalent or open. If it is key-equivalent, we have already observed that there must exist some key $`𝚔^{}`$ such that $`𝚙𝚔^{}S^{}`$, so $`I(𝚙)\mathrm{}`$. If it is open, suppose it is in equivalence class $`c`$. Then $`𝚙𝚔_cS^{}`$, since $`𝚙𝚙S^+`$ by (ClR). Again, it follows that $`I(𝚙)\mathrm{}`$. Conversely, suppose that $`I(𝚙)\mathrm{}`$. Thus, $`𝚙𝚔S^{}`$ for some key $`𝚔`$. If $`𝚙𝚔S^+`$, then by (ClK), $`𝚙\text{’s}𝚔_0𝚔_0S^+`$, so $`𝚙`$ is not empty. If $`𝚙𝚔S^+`$, then $`𝚔=𝚔_c`$, and there is some $`𝚚c`$ such that $`𝚙𝚚S^+`$. Since $`𝚚`$ is open, $`𝚚`$ cannot be empty, so $`𝚚\text{’s}𝚔_0𝚔_0S^+`$. Moreover, by (ClLM), $`𝚙\text{’s}𝚔_0𝚚\text{’s}𝚔_0S^+`$. Thus, by (ClT), $`𝚙\text{’s}𝚔_0𝚔_0S^+`$, so $`𝚙`$ is nonempty. ###### Lemma A.7 : For all expressions $`𝚙P`$, we have $`[[𝚙]]_{w,l,𝚔_0}=I(𝚙)`$. Proof: We proceed by induction on $`|𝚙|`$ (as defined in Lemma A.4). The claim is immediate from the definitions in case $`𝚙`$ is a global name or a local name. Suppose that $`𝚙`$ is a key $`𝚔_1`$. Then $`[[𝚙]]_{w,l,𝚔_0}=\{𝚔_1\}`$. Since $`𝚔_1𝚔_1S^{}`$ by construction, it follows that $`𝚔_1I(𝚔_1)`$. It remains to show that $`I(𝚔_1)\{𝚔_1\}`$. Suppose $`(𝚔_1𝚔)S^{}`$. By Lemma A.5, we cannot have $`𝚔=𝚔_0`$. Since $`S^+`$ is $`\text{AX}_{\mathrm{𝑖𝑛𝑓}}`$-consistent and closed under (ClKD), if $`𝚔P`$ we must have $`𝚔_1=𝚔`$. The remaining possibility for $`𝚔`$, that it equals $`𝚔_c`$ for some $`cO`$, cannot happen. For if so, only the second clause of the definition of $`S^{}`$ could explain $`(𝚔_1𝚔)S^{}`$. But then we have $`(𝚔_1𝚚)S^+`$ for some $`𝚚c`$. This contradicts the assumption that $`c`$ is an equivalence class of open expressions. Finally, suppose that $`|𝚙|>1`$. Let $`𝚙^{}`$ be the left-associative variant of $`𝚙`$. It is clear from the semantics that $`[[𝚙]]_{w,l,𝚔_0}=[[𝚙^{}]]_{w,l,𝚔_0}`$. Morover, (ClLV) and (ClT) guarantee that $`I(𝚙)=I(𝚙^{})`$. Thus, it suffices to prove that $`I(𝚙^{})=[[𝚙^{}]]_{w,l,𝚔_0}`$. Suppose that $`𝚙^{}=𝚚\text{’s}𝚛`$. The definition of length guarantees that $`|𝚙^{}|=|𝚙|>|𝚚|`$, so the induction hypothesis applies to $`𝚚`$. Since $`𝚙^{}`$ is associated to the left, $`𝚛GKN`$. Suppose that $`𝚛=𝚐GK`$. Note that $`[[𝚚\text{’s}𝚐]]_{w,l,𝚔_0}=\mathrm{}`$ if $`[[𝚚]]_{w,l,𝚔_0}=\mathrm{}`$ and $`[[𝚚\text{’s}𝚐]]_{w,l,𝚔_0}=[[𝚐]]_{w,l,𝚔_0}`$ if $`[[𝚚]]_{w,l,𝚔_0}\mathrm{}`$. We consider these two cases separately. Suppose first that $`[[𝚚]]_{w,l,𝚔_0}=\mathrm{}`$, so $`[[𝚙^{}]]_{w,l,𝚔_0}=\mathrm{}`$. By the induction hypothesis, $`I(𝚚)=\mathrm{}`$. To show that $`I(𝚙^{})=\mathrm{}`$, we show that $`𝚙^{}`$ is empty. Suppose not. Then $`(𝚙^{})\text{’s}𝚔_0𝚔_0S^+`$. Since $`S^+`$ contains either $`𝚚\text{’s}𝚔_0𝚔_0`$ or $`\neg (𝚚\text{’s}𝚔_0𝚔_0)`$ and $`S^+`$ is $`\text{AX}_{\mathrm{𝑖𝑛𝑓}}`$-consistent, by Nonemptiness(c), Associativity, and Transitivity, we must have $`𝚚\text{’s}𝚔_0𝚔_0S^+`$. Thus, $`𝚚`$ is not empty. By Lemma A.6, $`I(𝚚)\mathrm{}`$, a contradiction. Hence, $`𝚙^{}`$ is empty. It now follows from Lemma A.6 that $`I(𝚙^{})=\mathrm{}`$, as desired. Consider next the case where $`[[𝚚]]_{w,l,𝚔_0}\mathrm{}`$, so $`[[𝚙^{}]]_{w,l,𝚔_0}=[[𝚚\text{’s}𝚐]]_{w,l,𝚔_0}=[[𝚐]]_{w,l,𝚔_0}`$. To show that $`[[𝚙^{}]]_{w,l,𝚔_0}=I(𝚙^{})`$, we show that $`I(𝚙^{})=I(𝚐)`$. The result then follows from the induction hypothesis. By the induction hypothesis, $`I(𝚚)\mathrm{}`$, so by Lemma A.6, $`𝚚`$ is not empty. It follows from (ClG) that $`𝚚\text{’s}𝚐𝚐S^+`$. Suppose that $`𝚔I(𝚐)`$. If $`𝚔P_1`$, then $`𝚐𝚔S^+`$, so by (ClT), $`𝚚\text{’s}𝚐𝚔S^+`$ and $`𝚔I(𝚙^{})`$. If $`𝚔=𝚔_c`$ for some $`cO`$, then $`𝚐𝚚^{}S^+`$ for some $`𝚚^{}c`$. Thus, $`𝚙^{}𝚚^{}S^+`$ by (ClT) and we obtain that $`𝚙^{}𝚔S^{}`$ by construction of $`S^{}`$. Thus, $`I(𝚐)I(𝚙^{})`$. For the opposite containment, note that by (ClCG) we have $`𝚐𝚚\text{’s}𝚐S^+`$. Arguing as above, we obtain using (ClT) that $`I(𝚐)I(𝚙^{})`$. This completes the proof that $`I(𝚙^{})=I(𝚐)`$. It remains to deal with the case that $`𝚙^{}`$ has of the form $`𝚚\text{’s}𝚗`$, where $`𝚗`$ is a local name. There are three possibilities: $`𝚚`$ is empty, key-equivalent or open. If $`𝚚`$ is empty, then by Lemma A.6 and the induction hypothesis, $`I(𝚚)=\mathrm{}`$ and $`[[𝚚]]_{w,l,𝚔_0}=\mathrm{}`$. It follows that $`[[𝚙^{}]]_{w,l,𝚔_0}=\mathrm{}`$. Moreover, using Nonemptiness(c), Associativity, and Transitivity as above, it follows that $`𝚙^{}`$ is empty and hence by Lemma A.6, $`I(𝚙^{})=\mathrm{}`$, as desired. If $`𝚚`$ is key-equivalent, say $`𝚚𝚔_1`$, then $`𝚚𝚔_1S^+`$ and $`𝚔_1𝚚S^+`$. Using Key Distinctness and the consistency of $`S^+`$, it easily follows that $`I(𝚚)=\{𝚔_1\}`$. By the induction hypothesis, $`[[𝚚]]_{w,l,𝚔_0}=\{k_1\}`$. Thus, $`[[𝚙^{}]]_{w,l,𝚔_0}=l(𝚔_1,𝚗)`$. By construction, $`l(𝚔_1,𝚗)=I(𝚔_1\text{’s}𝚗)=I(𝚙^{})`$, as desired. Finally, suppose that $`𝚚`$ is open. If $`𝚔I(𝚙^{})`$, then it is immediate from the construction that that $`𝚚𝚔_{[𝚚]}S^{}`$ and $`𝚔l(𝚔_{[𝚚]},𝚗)`$. By the induction hypothesis, $`𝚔_{[𝚚]}[[𝚚]]_{w,l,𝚔_0}`$, so $`𝚔[[𝚙^{}]]_{w,l,𝚔_0}=_{𝚔^{}[[𝚚]]_{w,l,𝚔_0}}l(𝚔^{},𝚗)`$. Thus, $`I(𝚙^{})[[𝚙^{}]]_{w,l,𝚔_0}`$ if $`𝚙^{}`$ is open. For the opposite containment, suppose that $`𝚔[[𝚙^{}]]_{w,l,𝚔_0}`$. This means that there is some key $`𝚔^{}`$ such that $`𝚔^{}[[𝚚]]_{w,l,𝚔_0}`$ and $`𝚔l(𝚔^{},𝚗)`$. By the induction hypothesis, $`𝚔^{}I(𝚚)`$, so $`𝚚𝚔^{}S^{}`$. If $`𝚔^{}P_1`$, then $`𝚚𝚔^{}S^+`$ and $`(𝚔^{})\text{’s}𝚗𝚔S^+`$. (Since $`𝚚P`$ and $`𝚚𝚔^{}S^{}`$, we cannot have $`𝚔^{}=𝚔_0`$, by Lemma A.5.) By (ClLM), $`𝚚\text{’s}𝚗(𝚔^{})\text{’s}𝚗S^+`$, so by (ClT) we get $`𝚚\text{’s}𝚗𝚔S^+`$. Hence, $`𝚔I(𝚙^{})`$. If $`𝚔^{}=𝚔_c`$, where $`c`$ is an open equivalence class, then from $`𝚚𝚔^{}S^{}`$ it follows that $`𝚚𝚚^{}S^+`$ for some $`𝚚^{}c`$. From $`𝚔l(𝚔_c,𝚗)`$ it follows that $`(𝚛^{})\text{’s}𝚗𝚔S^{}`$ for some $`𝚛^{}c`$. By construction of $`S^+`$ we must have $`(𝚛^{})\text{’s}𝚗P_1`$, and since $`𝚛^{}𝚚^{}`$, we have $`𝚚^{}𝚛^{}S^+`$. By (ClT) we obtain $`𝚚𝚛^{}S^+`$, and hence by (ClLM) that $`𝚚\text{’s}𝚗(𝚛^{})\text{’s}𝚗S^+`$. Now notice that it follows from $`𝚚\text{’s}𝚗(𝚛^{})\text{’s}𝚗S^+`$ and $`(𝚛^{})\text{’s}𝚗𝚔S^{}`$ that $`𝚚\text{’s}𝚗𝚔S^{}`$. If $`𝚔P_1`$, this is immediate from (ClT). In case $`𝚔=𝚔_d`$ for some open class $`d`$, we have $`(𝚛^{})\text{’s}𝚗𝚝S^+`$ for some $`𝚝d`$. But then $`𝚚\text{’s}𝚗𝚝S^+`$ by (ClT); by definition of $`S^{}`$ we get that $`𝚚\text{’s}𝚗𝚔S^{}`$. This completes the proof. ###### Lemma A.8 : For all formulas $`\psi \text{Sub}(\varphi )E`$, we have $`\psi S`$ iff $`w,l,𝚔_0_\mathrm{o}\psi `$. Proof: We first show that by induction on the structure of $`\psi \text{Sub}(\varphi )E`$ that $`\psi S`$ iff $`w,l,𝚔_0\psi `$, and then show that the assignment $`l`$ is consistent with $`w`$. It is immediate from the construction of $`w`$ that $`w,l,𝚔_0\psi `$ iff $`\psi S`$ for $`\psi `$ of the form $`𝚔\mathrm{𝑐𝑒𝑟𝑡}(𝚗𝚙)`$. If $`\psi `$ has the form $`𝚙𝚚`$, note that $`w,l,𝚔_0𝚙𝚚`$ iff $`[[𝚙]]_{w,l,𝚔_0}[[𝚚]]_{w,l,𝚔_0}`$ iff (by Lemma A.7) iff $`I(𝚙)I(𝚚)`$. Thus, it suffices to show that $`I(𝚙)I(𝚚)`$ iff $`𝚙𝚚S^+`$, for $`𝚙,𝚚P`$. The “if” direction is immediate from (ClT): If $`𝚔I(𝚚)`$ then $`𝚚𝚔S^{}`$, so by (ClT) and the construction of $`S^{}`$, $`𝚙𝚔S^{}`$ and thus $`𝚔I(𝚙)`$. For the “only if” direction, suppose by way of contradiction that $`I(𝚙)I(𝚚)`$ but $`𝚙𝚚S^+`$. Then, by construction, $`\neg (𝚙𝚚)S^+`$. We consider three cases, depending on whether $`𝚚`$ is empty, key-equivalent, or open. Note first that $`𝚚`$ cannot be empty: $`\neg (𝚙𝚚)S^+`$, so by (ClN) we have $`𝚚\text{’s}𝚔_0𝚔_0S^+`$. Suppose that $`𝚚`$ is key-equivalent, with $`𝚔𝚚S^+`$. If $`𝚙𝚔S^+`$ then, by (ClT), $`𝚙𝚚S^+`$, but this is not possible because $`S^+`$ is $`\text{AX}_{\mathrm{𝑖𝑛𝑓}}`$-consistent. Thus $`𝚙𝚔S^+`$. Since $`𝚔P`$, $`𝚙𝚔S^{}`$, and thus $`𝚔I(𝚙)I(𝚚)`$, giving us the desired contradiction. Finally, suppose $`𝚚`$ is open. By construction, $`𝚚𝚔_{[𝚚]}S^{}`$. Moreover, we cannot have $`𝚙𝚔_{[𝚚]}S^{}`$, for then there would exist $`𝚛𝚚`$ such that $`𝚙𝚛S^+`$. Using (ClT), it would follow that $`𝚙𝚚S^+`$, which is impossible since $`S^+`$ is $`\text{AX}_{\mathrm{𝑖𝑛𝑓}}`$-consistent. Thus, $`𝚔_{[𝚚]}I(𝚙)I(𝚚)`$, giving the required contradiction, and completing the proof in the case that $`\psi `$ is of the form $`𝚙𝚚`$. If $`\psi `$ is of the form $`\neg \psi ^{}`$ or $`\psi _1\psi _2`$, the result is immediate from the induction hypothesis (in the latter case, we need the fact that if $`\psi _1\psi _2\text{Sub}(\varphi )E`$, then in fact $`\psi \psi _2\text{Sub}(\varphi )`$, so $`\psi _1,\psi _2\text{Sub}(\varphi )`$ and the induction hypothesis applies). This completes the induction proof. To show that the assignment $`l`$ is consistent with $`w`$, suppose that $`𝚗𝚙c(𝚔)`$. Then, by construction, $`𝚔\mathrm{𝑐𝑒𝑟𝑡}(𝚗𝚙)S`$. By (ClKL), we have $`𝚔\text{’s}𝚗𝚔\text{’s}𝚙S^+`$. By what we have just shown $`w,l,𝚔_0𝚔\text{’s}𝚗𝚔\text{’s}𝚙`$. It follows that $`w,l,𝚔𝚗𝚙`$. Thus, $`l`$ is consistent with $`w`$. Thus, we have shown that $`\varphi `$ is satisfiable, completing the proof of Theorem 3.2 in the case that $`K`$ is infinite. The same argument works without change if $`K`$ is finite but $`|K|2|\varphi |^2`$. (A consequence of this is that we do not need to use the axioms Witnesses and Current Principal to derive a valid formula $`\varphi `$ in $`\text{AX}_{\mathrm{𝑓𝑖𝑛}}`$ if $`2|\varphi |^2|K|`$.) Moreover, the proof shows that Proposition 3.3 holds if $`|K|2|\varphi |^2`$. Now suppose that $`K2|\varphi |^2`$. We show that if $`\varphi `$ is $`\text{AX}_{\mathrm{𝑓𝑖𝑛}}`$-consistent, then $`\varphi `$ is satisfiable. The proof is in the spirit of that in the case of $`\text{AX}_{\mathrm{𝑖𝑛𝑓}}`$, but simpler. Now let $`P`$ be the least set of principal expressions containing all principal expressions that appear in $`\varphi `$ and closed under subexpressions. Let $`F`$ consist of all formulas of the form $`𝚙𝚔^{}`$ and $`𝚔\text{’s}𝚙𝚔^{}`$, where $`𝚙P`$ and $`𝚔,𝚔^{}K`$. Let $`S`$ be an $`\text{AX}_{\mathrm{𝑓𝑖𝑛}}`$-consistent set containing $`\varphi `$ and, for every formula $`\psi \text{Sub}(\varphi )F`$, either $`\psi `$ or $`\neg \psi `$. Since $`\varphi `$ is $`\text{AX}_{\mathrm{𝑓𝑖𝑛}}`$-consistent, there must be some $`\text{AX}_{\mathrm{𝑓𝑖𝑛}}`$-consistent set $`S`$ of this form. There must be some key $`𝚔_0K`$ such that for every local name in $`P`$ and key $`𝚔K`$, we have $`𝚗𝚔S`$ iff $`𝚔_0\text{’s}𝚗𝚔S`$. For otherwise, for each key $`𝚔`$, there is some local name $`𝚗_𝚔`$ and key $`𝚔_𝚔`$ such that either both $`𝚗_𝚔𝚔_𝚔`$ and $`\neg (𝚔\text{’s}𝚗_𝚔𝚔_𝚔)`$ are in $`S`$ or both $`\neg (𝚗_𝚔𝚔_𝚔)`$ and $`𝚔\text{’s}𝚗_𝚔𝚔_𝚔`$ are in $`S`$. This means that $`S`$ is inconsistent with the axiom Current Principal. Define the local assignment $`l`$ so that $`l(𝚔,𝚗)=\{𝚔^{}:𝚔\text{’s}𝚗𝚔^{}S\}`$. Similar to the case for $`\text{AX}_{\mathrm{𝑖𝑛𝑓}}`$, define the world $`w=(\beta ,c)`$ by taking $`\beta (𝚐)=\{𝚔K|𝚐𝚔S^+\}`$ and defining $`c(𝚔)`$, for each key $`𝚔`$, to be the set of formulas $`𝚗𝚙`$ such that $`𝚔\mathrm{𝑐𝑒𝑟𝑡}(𝚗𝚙)S`$. Now we have the following analogue to Lemma A.8. ###### Lemma A.9 : For all formulas $`\psi \text{Sub}(\varphi )F`$, we have $`\psi S`$ iff $`w,l,𝚔_0_\mathrm{o}\psi `$. Proof: Again we first show that by induction on the structure of $`\psi \text{Sub}(\varphi )E`$ that $`\psi S`$ iff $`w,l,𝚔_0\psi `$, and then show that the assignment $`l`$ is consistent with $`w`$. It is immediate from the construction of $`w`$ that $`w,l,𝚔_0\psi `$ iff $`\psi S`$ for $`\psi `$ of the form $`𝚔\mathrm{𝑐𝑒𝑟𝑡}(𝚗𝚙)`$. We next show that the result holds if $`\psi `$ is of the form $`𝚙𝚔^{}`$, for $`𝚙P`$, by induction on the structure of $`𝚙`$. We strengthen the induction hypothesis to also show that $`w,l,𝚔_0𝚔\text{’s}𝚙𝚔^{}`$ iff $`𝚔\text{’s}𝚙𝚔S`$. If $`𝚙`$ is a key $`𝚔_1`$, then $`w,l,𝚔_0𝚔_1𝚔^{}`$ iff $`𝚔^{}=𝚔_1`$ and by Reflexivity and Key Distinctness, $`𝚔_1𝚔^{}S`$ iff $`𝚔_1=𝚔^{}`$. Similarly, $`w,l,𝚔_0𝚔\text{’s}𝚔_1𝚔^{}`$ iff $`w,l,𝚔_0𝚔_1𝚔^{}`$ iff $`𝚔_1𝚔^{}S`$ iff $`𝚔\text{’s}𝚔_1𝚔^{}S`$, by Transitivity, Key Globality, and Converse of Globality (using the fact that $`S`$ is $`\text{AX}_{\mathrm{𝑓𝑖𝑛}}`$-consistent). If $`𝚙`$ is a global identifier $`𝚐`$, $`w,l,𝚔_0𝚐𝚔^{}`$ iff $`𝚐𝚔^{}S`$ by the definition of $`\beta `$. The argument for $`𝚔\text{’s}𝚐𝚔^{}`$ is identical to the case that $`𝚙=𝚔`$. If $`𝚙`$ is the local name $`𝚗`$, then $`w,l,𝚔_0𝚗𝚔^{}`$ iff $`𝚔^{}l(𝚔_0,𝚗)`$ iff $`𝚔_0\text{’s}𝚗𝚔^{}S`$ iff $`𝚗𝚔^{}S`$, by choice of $`𝚔_0`$. Similarly, $`w,l,𝚔_0𝚔\text{’s}𝚗𝚔^{}`$ iff $`𝚔^{}l(𝚗,𝚔)`$ iff $`𝚔\text{’s}𝚗𝚔^{}S`$. Finally, if $`𝚙`$ is of the form $`𝚚\text{’s}𝚛`$, then $`w,l,𝚔_0𝚚\text{’s}𝚛𝚔^{}`$ iff there exists a key $`𝚔^{\prime \prime }`$ such that $`w,l,𝚔_0𝚚𝚔^{\prime \prime }`$ and $`w,l,𝚔_0(𝚔^{\prime \prime })\text{’s}𝚛𝚔^{}`$ iff (by the induction hypothesis) there exists a key $`𝚔^{\prime \prime }`$ such that $`𝚚𝚔^{\prime \prime }S`$ and $`(𝚔^{\prime \prime })\text{’s}𝚛𝚔^{}S`$ iff $`𝚚\text{’s}𝚛𝚔^{}S`$. The “only if” direction of the last equivalence follows using Left Monotonocity and Transitivity; the “if” direction follows from Witnesses. The argument for $`𝚔\text{’s}(𝚚\text{’s}𝚛)𝚔^{}`$ is identical, using Associativity: $`w,l,𝚔_0𝚔\text{’s}(𝚚\text{’s}𝚛)𝚔^{}`$ iff there exists a key $`𝚔^{\prime \prime }`$ such that $`w,l,𝚔_0𝚔\text{’s}𝚚𝚔^{\prime \prime }`$ and $`w,l,𝚔_0(𝚔^{\prime \prime })\text{’s}𝚛𝚔^{}`$ iff there exists a key $`𝚔^{\prime \prime }`$ such that $`𝚔\text{’s}𝚚𝚔^{\prime \prime }S`$ and $`(𝚔^{\prime \prime })\text{’s}𝚛𝚔^{}S`$ iff $`𝚔\text{’s}(𝚚\text{’s}𝚛)𝚔^{}S`$. We now continue with our induction in the case that $`𝚙𝚚`$. Note that $`w,l,𝚔_0𝚙𝚚`$ iff $`w,l,𝚔_0𝚚𝚔^{}`$ implies $`w,l,𝚔_0𝚙𝚔^{}`$ for all $`𝚔^{}K`$ iff (by the induction hypothesis) $`𝚚𝚔^{}S`$ implies $`𝚙𝚔^{}S`$ iff $`𝚙𝚚S`$. The “only if” direction of the last equivalence follows immediately from Transitivity; the “if” direction follows from Witnesses. We complete the induction proof by observing that if $`\psi `$ is of the form $`\neg \psi `$ or $`\psi _1\psi _2`$, the result follows immediately from the induction hypothesis. To show that $`l`$ is consistent with $`w`$, suppose that $`𝚗𝚙c(𝚔)`$. By construction, this means that $`𝚔\mathrm{𝑐𝑒𝑟𝑡}(𝚗𝚙)S`$. By Key Linking, we must also have $`𝚔\text{’s}𝚗𝚔\text{’s}𝚙S`$. By what we have just shown, $`w,l,𝚔_0𝚔\text{’s}𝚗𝚔\text{’s}𝚙`$. It follows that $`w,l,𝚔𝚗𝚙`$. Thus, $`l`$ is consistent with $`w`$. This completes the proof of Theorem 3.2 in the case that $`K`$ is finite. Note that since we can assume without loss of generality that $`|K|2|\varphi |^2`$ here (otherwise the argument for the case that $`K`$ is infinite applies) the proof also shows that Proposition 3.3 holds. Theorem 3.5: The same formulas are c-valid and o-valid; i.e., for all formulas $`\varphi `$, we have $`_\mathrm{o}\varphi `$ iff $`_\mathrm{c}\varphi `$. Proof: We show that $`\neg \varphi `$ is o-satisfiable iff $`\neg \varphi `$ is c-satisfiable, which is equivalent to the claim. The direction from c-satisfiability to o-satisfiability is straightforward: Since for every world $`w`$ the local name assignment $`l_w`$ is $`w`$-consistent, it follows from $`w,𝚔_\mathrm{c}\neg \varphi `$ that $`w,l_w,𝚔_\mathrm{o}\neg \varphi `$. Thus, it remains to show that if $`\neg \varphi `$ is o-satisfiable, then it is c-satisfiable. So suppose that $`\neg \varphi `$ is o-satisfiable. By Proposition 3.3, there is a world $`w=(\beta ,c)`$, local name assignment $`l`$, and principal $`𝚔`$ such that $`w,l,𝚔_o\neg \varphi `$ and a finite subset $`K^{}`$ of $`K`$ such that $`l(𝚔^{},𝚗)K^{}`$ for all $`𝚔^{}K`$ and $`𝚗N`$, and $`\beta (𝚐)K^{}`$ for all global names $`𝚐`$. By standard propositional reasoning, $`\neg \varphi `$ is equivalent to a disjunctive normal form expression in which the atoms are of the form $`𝚙𝚚`$ and $`𝚔_1\mathrm{𝑐𝑒𝑟𝑡}\psi `$, where $`𝚙`$ and $`𝚚`$ are principal expressions, $`𝚔_1`$ is a key, and $`\psi `$ is a formula. If $`w,l,𝚔_\mathrm{o}\neg \varphi `$ then one of the disjuncts $`\sigma `$ is satisfied, i.e., $`w,l,𝚔_\mathrm{o}\sigma `$. Suppose that $`\sigma `$ is the conjunction of the formulas in the set $`AB`$, where 1. $`A`$ is a set of formulas of the form $`𝚙𝚚`$ or $`\neg (𝚙𝚚)`$, 2. $`B`$ is a set of formulas of the form $`𝚔_1\mathrm{𝑐𝑒𝑟𝑡}\psi `$ or $`\neg (𝚔_1\mathrm{𝑐𝑒𝑟𝑡}\psi )`$. Let $`K_\varphi `$ be the set of keys that appear in the formula $`\varphi `$ together with $`K^{}`$ and $`𝚔`$. Let $`N_\varphi `$ be the set of local names that appear in $`\varphi `$. Define the world $`w^{}=(\beta ^{},c^{})`$ as follows. Take the interpretation of global names $`\beta ^{}`$ to be equal to $`\beta `$, the interpretation of global names in $`w`$. Define $`c^{}`$ by taking the set of certificates $`c^{}(𝚔^{})`$ to be the empty if $`𝚔^{}K_\varphi `$ and to consist of $`c(𝚔^{})`$ together with all certificates of the form $`𝚗𝚙_{𝚔^{\prime \prime },\varphi }`$ if $`𝚔^{}K_\varphi `$, $`𝚗N_\varphi `$, and $`𝚔^{\prime \prime }l(𝚗,𝚔^{})`$, where $`𝚙_{𝚔^{\prime \prime },\varphi }`$ is a principal expression of the form $`(𝚔^{\prime \prime })\text{’s}(𝚔^{\prime \prime })\text{’s}\mathrm{}(𝚔^{\prime \prime })`$ that does not appear in $`\varphi `$. (Clearly we can make the expression sufficiently long so as to ensure it does not appear in $`\varphi `$.) Clearly $`_{𝚔^{}K}c(𝚔^{})`$ is finite. We show that $`w^{},𝚔_\mathrm{c}\sigma `$. It follows from this that $`w^{},𝚔_\mathrm{c}\neg \varphi `$. Note first that from the fact that $`c(𝚔^{})c^{}(𝚔^{})`$ for all $`𝚔^{}`$, it follows that $`w^{},𝚔_\mathrm{c}𝚔^{}\mathrm{𝑐𝑒𝑟𝑡}\psi `$ for all formulas $`𝚔^{}\mathrm{𝑐𝑒𝑟𝑡}\psi `$ in $`B`$. Moreover, if $`\neg (𝚔^{}\mathrm{𝑐𝑒𝑟𝑡}\psi )`$ is in $`B`$ then, since the expressions $`𝚙_{𝚔^{\prime \prime },\varphi }`$ on the right-hand side of the certificates in $`c^{}(𝚔^{})c(𝚔)`$ do not appear in $`\varphi `$ it follows that $`w^{},𝚔_\mathrm{c}\neg (𝚔^{}\mathrm{𝑐𝑒𝑟𝑡}\psi )`$. Thus $`w^{},𝚔_\mathrm{c}B`$. It remains to show that the formulas in $`A`$ are satisfied. To show this, we show that $`l_w^{}(𝚗,𝚔^{})=l(𝚗,𝚔^{})`$ for all $`𝚗N_\varphi `$ and $`𝚔^{}K_\varphi `$. (2) It easily follows from (2), the fact that all keys in $`\varphi `$ are in $`K^{}`$, and the fact that global names have the same interpretation in $`w`$ and $`w^{}`$ that $`[[p]]_{w^{},l_w^{},𝚔^{}}=[[p]]_{w,l,𝚔^{}}`$ for all principal expressions $`p`$ occurring in $`A`$ and all keys $`𝚔^{}K_\varphi `$. This in turn is easily seen to imply that $`w^{},𝚔_\mathrm{c}A`$. It remains to prove (2). It is almost immediate from the definition of $`l^{}`$ that $`l_w^{}(𝚗,𝚔^{})l(𝚗,𝚔^{})`$ for all $`𝚗N_\varphi `$ and $`𝚔^{}K_\varphi `$. For the opposite containment, we prove by induction on $`j`$ that $`(T_w^{}j)(𝚗,𝚔^{})l(𝚗,𝚔^{})`$ for all $`j𝐍`$, $`𝚗N_\varphi `$, and $`𝚔^{}K_\varphi `$. The base case $`j=0`$ is trivial. For the induction step, suppose that $`j=j^{}+1`$ and $`𝚔^{\prime \prime }(T_w^{}j)(𝚗,𝚔^{})`$. Thus, $`𝚔^{\prime \prime }(T_w^{}(T_w^{}j^{}))(𝚗,𝚔^{})`$, which means that $`𝚔^{\prime \prime }[[𝚙]]_{w^{},T_w^{}j^{},𝚔^{}}`$ for some principal expression $`𝚙`$ such that $`𝚗𝚙c^{}(𝚔^{})`$. There are two possibilities: (1) $`𝚗𝚙c(𝚔^{})`$ or (2) $`𝚗𝚙c^{}(𝚔^{})c(𝚔^{})`$. In case (2), $`𝚙`$ must be of the form $`𝚙_{𝚔_1,\varphi }`$ so $`[[𝚙]]_{w^{},T_w^{}j^{},𝚔^{}}=\{𝚔_1\}`$ and $`𝚔_1=𝚔^{\prime \prime }`$. But in this case, by construction, $`𝚔^{\prime \prime }l(𝚗,𝚔^{})`$. In case (1), using the induction hypothesis and the fact that global names and keys in $`𝚙`$ have the same interpretation in $`w`$ and $`w^{}`$ (this interpretation being a subset of $`K^{}`$), we get that $`[[𝚙]]_{w^{},T_w^{}j^{},𝚔^{}}[[𝚙]]_{w,l,𝚔^{}}`$. Thus, $`𝚔^{\prime \prime }[[𝚙]]_{w,l,𝚔^{}}`$. Because $`l`$ is $`w`$-consistent and $`𝚗𝚙c(𝚔^{})`$, we again obtain that $`𝚔^{\prime \prime }l(𝚗,𝚔^{})`$, as required. Since $`l_w^{}(𝚗,𝚔^{})`$ is the union of the $`(T_w^{}j)(𝚗,𝚔^{})`$, it follows that $`l_w^{}(𝚗,𝚔^{})=l(𝚗,𝚔^{})`$. This completes the proof of (2). Proposition 3.8: Let $`\mathrm{\Gamma }`$ be any c-satisfiable boolean combination of formulas of the form $`𝚔\mathrm{𝑐𝑒𝑟𝑡}\varphi `$, and let $`\mathrm{\Delta }`$ be any boolean combination of formulas of the form $`𝚙𝚚`$ where neither $`𝚙`$ nor $`𝚚`$ contains a local name. Then $`_\mathrm{c}\mathrm{\Gamma }\mathrm{\Delta }`$ iff $`_\mathrm{c}\mathrm{\Delta }`$. Proof: Clearly $`_\mathrm{c}\mathrm{\Delta }`$ implies $`_\mathrm{c}\mathrm{\Gamma }\mathrm{\Delta }`$. For the converse, suppose by way of contradiction that $`_\mathrm{c}\mathrm{\Gamma }\mathrm{\Delta }`$ and there is a world $`w=(\beta ,c)`$ and a principal $`𝚔`$ such that $`w,𝚔_\mathrm{c}\neg \mathrm{\Delta }`$. Since $`\mathrm{\Gamma }`$ is assumed to be c-satisfiable, there exists a world $`w^{}=(\beta ^{},c^{})`$ and a principal $`𝚔^{}`$ such that $`w^{},𝚔^{}_\mathrm{c}\mathrm{\Gamma }`$. Let $`w^{\prime \prime }`$ be the world $`(\beta ,c^{})`$. Then a straightforward induction shows that for all principal expressions $`𝚙`$ not containing a local name, we have $`[[𝚙]]_{w^{\prime \prime },l_{w^{\prime \prime }},𝚔}=[[𝚙]]_{w,l_w,𝚔}`$. Moreover, for all keys $`𝚔_1`$ and formulas $`\varphi `$, we have $`w^{\prime \prime },𝚔_\mathrm{c}𝚔_1\mathrm{𝑐𝑒𝑟𝑡}\varphi `$ iff $`w^{},𝚔^{}_\mathrm{c}𝚔_1\mathrm{𝑐𝑒𝑟𝑡}\varphi `$. It follows that $`w^{\prime \prime },𝚔_\mathrm{c}\mathrm{\Gamma }\neg \mathrm{\Delta }`$, giving us our desired contradiction. Theorem 4.1: Suppose $`𝚔_1,𝚔_2`$ are principals, $`w=(\beta ,c)`$ is a world, and $`𝚙`$ is a principal expression. Let $`E_w`$ be the set of all the formulas $`𝚐𝚔`$ for all global names $`𝚐`$ and keys $`𝚔\beta (𝚐)`$ and the formulas $`𝚔\mathrm{𝑐𝑒𝑟𝑡}\varphi `$ for all keys $`𝚔`$ and formulas $`\varphi c(𝚔)`$. The following are equivalent: 1. $`𝚔_1\text{REF2}(𝚔_2,\beta ,c,𝚙)`$, 2. $`w,𝚔_2_\mathrm{c}𝚙𝚔_1`$, 3. $`w^{},𝚔_2_\mathrm{c}𝚙𝚔_1`$ for all worlds $`w^{}w`$, 4. $`E_w_c𝚔_2\text{’s}𝚙𝚔_1`$, 5. $`E_w_o𝚔_2\text{’s}𝚙𝚔_1`$. Proof: The presentation of REF2 in Figure 1 is still slightly informal, combining recursion and nondeterminism. To make it fully precise, define a computation tree of REF2 to be a finite tree labelled by expressions of the form “$`𝚔_1\text{REF2}(𝚔_2,\beta ,c,𝚙)`$”, such that if $`N`$ is a node so labelled, then one of the following four conditions holds: 1. $`𝚙`$ is a key $`𝚔`$, we have $`𝚔=𝚔_1=𝚔_2`$, and $`N`$ is a leaf of the tree, 2. $`𝚙`$ is a global name $`𝚐`$ and $`𝚔_1\beta (𝚐)`$, 3. $`𝚙`$ is a local name $`𝚗`$ and $`c(𝚔_2)`$ contains a formula $`𝚗𝚚`$ and $`N`$ has exactly one child, labelled “$`𝚔_1\text{REF2}(𝚔_2,\beta ,c,𝚚)`$”, 4. $`𝚙`$ is of the form $`𝚚\text{’s}𝚛`$ and $`N`$ has exactly two children, labelled “$`𝚔\text{REF2}(𝚔_2,\beta ,c,𝚚)`$” and “$`𝚔_1\text{REF2}(𝚔,\beta ,c,𝚛)`$”, for some key $`𝚔`$. We take $`𝚔_1\text{REF2}(𝚔_2,\beta ,c,𝚙)`$ to mean that there exists a computation tree of REF2 with root labelled “$`k_1\text{REF2}(𝚔_2,\beta ,c,𝚙)`$”. Given a world $`w=(\beta ,c)`$ and $`m𝐍`$, let $`l_m=T_wm`$. The following result establishes a correspondence between the stages of the computation of $`l_w`$ and the computation trees of REF2. The proof is by a straightforward induction on $`m`$, with a subinduction on the structure of $`𝚙`$. ###### Lemma A.10 : For all $`m𝐍`$, keys $`𝚔_1,𝚔_2`$, worlds $`w=(\beta ,c)`$, and principal expressions $`𝚙`$, we have $`𝚔_1[[𝚙]]_{w,l_m,𝚔_2}`$ iff there exists a computation tree of REF2 of height at most $`m`$ whose root is labelled $`𝚔_1\text{REF2}(𝚔_2,\beta ,c,𝚙)`$. Using the fact that $`l_w=\{l_m:m𝐍\}`$, Lemma A.1, and Lemma A.10, we obtain the equivalence between (1) and (2). The proof of the implication from (2) to (3) is by a straightforward induction on the structure of $`𝚙`$; that is, for fixed $`w^{}w`$, we show by induction on the structure of $`𝚙`$ that if $`w,𝚔_2_\mathrm{c}𝚙𝚔_1`$ then $`w^{},𝚔_2_\mathrm{c}𝚙𝚔_1`$. The opposite implication from (3) to (2) is trivial, since $`ww`$. For the implication from (3) to (4), suppose that (3) holds and (4) does not. Then for some world $`w^{}`$ and key $`𝚔`$ we have $`w^{},𝚔_\mathrm{c}E_w`$ and $`w^{},𝚔_\mathrm{c}\neg (𝚔_2\text{’s}𝚙𝚔_1)`$. The latter implies $`w^{},𝚔_2_\mathrm{c}\neg (𝚙𝚔_1)`$. Since $`w^{},𝚔_\mathrm{c}E_w`$, it follows that $`w^{}w`$. Thus, by (3), $`w^{},𝚔_2_\mathrm{c}𝚙𝚔_1`$, contradicting our assumption. The implication from (4) to (3) is immediate, since $`w^{},𝚔_2_\mathrm{c}E_w`$ for all $`w^{}w`$. Finally, the equivalence between (4) and (5) is just a special case of Theorem 3.5. Proposition 5.1: If $`M`$ represents $`w`$ and $`l`$ then for all principal expressions $`𝚙`$ and $`x,yKGN`$ we have $`M\tau _{x,y}(𝚙)`$ iff $`x,yK`$ and $`w,l,x𝚙y`$. Proof: By a straightforward induction on the structure of $`𝚙`$. The base cases, where $`𝚙KGN`$, are immediate from the definition of “represents” and the semantics of the logic. The inductive case, where $`𝚙=𝚚\text{’s}𝚛`$, is immediate from the semantics and the definition of the translation. Theorem 5.2: The minimal Herbrand model $`M_w`$ of $`\mathrm{\Sigma }_w`$ represents $`w`$ and $`l_w`$. Proof: (Sketch) The proof proceeds by showing a direct correspondence between the construction of the minimal Herbrand model of $`\mathrm{\Sigma }_w`$ and the fixpoint construction of $`l_w`$. The theory of logic programming \[Llo87\] associates with the Horn theory $`\mathrm{\Sigma }_w`$ an operator $`\mathrm{\Phi }_w`$ on the space of Herbrand models on the vocabulary $`V`$, defined by $`\mathrm{𝚗𝚊𝚖𝚎}(x,y,z)\mathrm{\Phi }_w(M)`$ if there exists a substitution instance of a formula in $`\mathrm{\Sigma }_w`$ of the form $`B\mathrm{𝚗𝚊𝚖𝚎}(x,y,z)`$ such that $`MB`$. The least Herbrand model $`M_w`$ of $`\mathrm{\Sigma }_w`$ is then equal to $`\mathrm{\Phi }_w\omega =_{m𝐍}\mathrm{\Phi }_wm`$, where $`\mathrm{\Phi }_w0=\mathrm{}`$ and $`\mathrm{\Phi }_wm+1=\mathrm{\Phi }_w(\mathrm{\Phi }_wm)`$ for $`m0`$. Let $`T_w`$ be the operator on local name assignments defined in the proof of Theorem 3.1. Using Proposition 5.1 to handle the rules in $`\mathrm{\Sigma }_w`$ corresponding to certificates, we may then show by a straightforward induction on $`m`$ that for all $`m1`$, the Herbrand model $`\mathrm{\Phi }m`$ represents the world $`w`$ and the local name assignment $`T_wm`$. It follows that $`M_w=\mathrm{\Phi }\omega `$ represents $`l_w=T_w\omega `$. Theorem 6.1: $`\text{AX}_{\mathrm{𝑖𝑛𝑓}}^{\mathrm{𝑠𝑒𝑙𝑓}}`$ (resp., $`\text{AX}_{\mathrm{𝑓𝑖𝑛}}^{\mathrm{𝑠𝑒𝑙𝑓}}`$) is a sound and complete axiomatization of LLNC<sup>s</sup> with respect to the open semantics if $`K`$ is infinite (resp., $`K`$ is finite). Proof: The argument is very similar to that in the proof of Theorem 3.2. First suppose that $`K`$ is infinite. We add the following clauses to the definition of $`P`$: 1. $`\mathrm{𝚂𝚎𝚕𝚏}P`$, 2. if $`𝚗P`$ is a local name then $`\mathrm{𝚂𝚎𝚕𝚏}\text{’s}𝚗P`$. We also add the following clauses to the definition of $`S^+`$, corresponding to the new axioms for $`\mathrm{𝚂𝚎𝚕𝚏}`$. 1. if $`\mathrm{𝚂𝚎𝚕𝚏}\text{’s}𝚙P`$ then $`\mathrm{𝚂𝚎𝚕𝚏}\text{’s}𝚙𝚙S^+`$ and $`𝚙\mathrm{𝚂𝚎𝚕𝚏}\text{’s}𝚙S^+`$, 2. if $`𝚙\text{’s}\mathrm{𝚂𝚎𝚕𝚏}P`$ then $`𝚙\text{’s}\mathrm{𝚂𝚎𝚕𝚏}𝚙S^+`$ and $`𝚙𝚙\text{’s}\mathrm{𝚂𝚎𝚕𝚏}S^+`$, 3. if $`\mathrm{𝚂𝚎𝚕𝚏}𝚙S^+`$ and $`𝚙\text{’s}𝚔𝚔S^+`$ then $`𝚙\mathrm{𝚂𝚎𝚕𝚏}S^+`$. Lemma A.5 still applies. The definitions following this lemma, up to and including that of $`S^{}`$ are unchanged. However, the construction of the model changes slightly. We no longer use $`𝚔_0`$ to represent the “current principal”, instead, we use the key $`𝚔_{}`$ that the construction associates with $`\mathrm{𝚂𝚎𝚕𝚏}`$. This could be either a key in $`P_1`$ or one of the keys $`𝚔_c`$ for $`cO`$, depending on whether $`\mathrm{𝚂𝚎𝚕𝚏}`$ is key-equivalent or open. Note that we cannot have $`\mathrm{𝚂𝚎𝚕𝚏}`$ empty (thanks to the Identity axiom). If $`\mathrm{𝚂𝚎𝚕𝚏}`$ is key-equivalent, then by (ClKD) it is equivalent to at most one key $`𝚔P`$. In this case, we define $`𝚔_{}=𝚔`$. If $`\mathrm{𝚂𝚎𝚕𝚏}`$ is open we define $`𝚔_{}`$ to be $`𝚔_c`$, where $`c=[\mathrm{𝚂𝚎𝚕𝚏}]`$. We now define $`w`$ and $`l`$ exactly as before, except that we now set $`l(𝚔_0,𝚗)=\mathrm{}`$, since we no longer use $`𝚔_0`$ as the “current principal.” The following lemma is the analogue of Lemma A.7. ###### Lemma A.11 : For all expressions $`𝚙P`$, we have $`[[𝚙]]_{w,l,𝚔_{}}=I(𝚙)`$. Proof: The proof is very similar to that of Lemma A.7; we just describe the modifications required. The base cases for $`𝚙`$ a global name or a key are identical. When $`𝚙=𝚗`$ is a local name, we proceed as follows. There are two possibilities, depending on whether $`𝚔_{}P`$ or not. Suppose first that $`𝚔_{}P`$. Then we have $`𝚔_{}\mathrm{𝚂𝚎𝚕𝚏}`$ and, by (ClLM) and (ClSP), $`𝚔_{}\text{’s}𝚗\mathrm{𝚂𝚎𝚕𝚏}\text{’s}𝚗𝚗`$. It then follows by (ClT) and construction of $`l`$ that $`𝚗𝚔S^{}`$ iff $`𝚔_{}\text{’s}𝚗𝚔S^{}`$ iff $`𝚔l(𝚔_{},𝚗)`$, as required. If $`𝚔_{}=𝚔_c`$ for $`c`$ an open class, we proceed as follows. If $`𝚔I(𝚗)`$, then we consider two cases, depending on whether $`𝚔P_1`$. If $`𝚔P_1`$, then $`𝚗𝚔S^+`$ and it follows that $`\mathrm{𝚂𝚎𝚕𝚏}\text{’s}𝚗𝚔`$ by (ClSP) and (ClT). Since $`\mathrm{𝚂𝚎𝚕𝚏}\mathrm{𝚂𝚎𝚕𝚏}`$ it is immediate that $`𝚔[[𝚙]]_{w,l,𝚔_{}}`$. Alternatively, if $`𝚔=𝚔_d`$, for $`dO`$, then we have $`𝚗𝚚S^+`$ for some $`𝚚d`$. By (ClSP) and (ClT) it follows that $`\mathrm{𝚂𝚎𝚕𝚏}\text{’s}𝚗𝚚S^+`$, hence $`\mathrm{𝚂𝚎𝚕𝚏}\text{’s}𝚗𝚔S^{}`$. As before, this implies that $`𝚔[[𝚗]]_{w,l,𝚔_{}}`$. For the opposite inclusion, suppose that $`𝚔[[𝚗]]_{w,l,𝚔_{}}`$. Since we are assuming that $`\mathrm{𝚂𝚎𝚕𝚏}`$ is open, there must be some $`𝚚\mathrm{𝚂𝚎𝚕𝚏}`$ such that $`𝚚\text{’s}𝚗𝚔S^{}`$. By (ClLM), we have $`\mathrm{𝚂𝚎𝚕𝚏}\text{’s}𝚗𝚚\text{’s}𝚗S^+`$. It follows using (ClT) that $`\mathrm{𝚂𝚎𝚕𝚏}\text{’s}𝚗𝚔S^{}`$, hence $`𝚗𝚔S^{}`$. This completes the argument for the base case of $`𝚗`$ a local name. There is now an additional base case for $`𝚙=\mathrm{𝚂𝚎𝚕𝚏}`$. Here, note that $`[[\mathrm{𝚂𝚎𝚕𝚏}]]_{w,l,𝚔_{}}=\{𝚔_{}\}`$. We therefore need to show that $`\mathrm{𝚂𝚎𝚕𝚏}𝚔S^{}`$ iff $`𝚔=𝚔_{}`$. When $`𝚔_{}P_1`$, we have $`\mathrm{𝚂𝚎𝚕𝚏}𝚔_{}`$, so $`\mathrm{𝚂𝚎𝚕𝚏}𝚔S^{}`$ iff $`𝚔_{}𝚔`$, and the claim follows by (ClKD) and (ClT) as in the base case for keys. The alternative is that $`𝚔_{}=𝚔_c`$ for $`c=[\mathrm{𝚂𝚎𝚕𝚏}]O`$. Since have $`\mathrm{𝚂𝚎𝚕𝚏}𝚔_cS^{}`$ by construction of $`S^{}`$, it remains to prove that if $`\mathrm{𝚂𝚎𝚕𝚏}𝚔S^{}`$ then $`𝚔=𝚔_c`$. Now we cannot have $`\mathrm{𝚂𝚎𝚕𝚏}𝚔S^{}`$ for $`𝚔P_1`$, for then by the argument above that $`\mathrm{𝚂𝚎𝚕𝚏}`$ is nonempty and (ClSE), we have $`𝚔\mathrm{𝚂𝚎𝚕𝚏}S^+`$, contradicting the assumption that $`c`$ is open. Thus, we must have $`𝚔=𝚔_d`$ for some $`dO`$. In this case, there exists $`𝚚d`$ such that $`\mathrm{𝚂𝚎𝚕𝚏}𝚚S^+`$. Since $`d`$ is open, we have $`𝚚\text{’s}𝚔_0𝚔_0S^+`$, hence $`𝚚\mathrm{𝚂𝚎𝚕𝚏}S^+`$ by (ClSE). Thus, $`\mathrm{𝚂𝚎𝚕𝚏}𝚚`$, and it follows that $`d=c`$, hence $`𝚔=𝚔_{}`$ as required. This completes the argument for the base case where $`𝚙=\mathrm{𝚂𝚎𝚕𝚏}`$. The inductive case is exactly as before, except that we need to consider the new case $`𝚙\text{’s}\mathrm{𝚂𝚎𝚕𝚏}`$. Here, we note that $`[[𝚙\text{’s}\mathrm{𝚂𝚎𝚕𝚏}]]_{w,l,𝚔_{}}=[[𝚙]]_{w,l,𝚔_{}}`$. Thus, by the induction hypothesis, we are required to prove that $`𝚙𝚔S^{}`$ iff $`𝚙\text{’s}\mathrm{𝚂𝚎𝚕𝚏}𝚔S^{}`$. This follows using (ClPS) and (ClT). The remainder of the proof in the case that $`K`$ is infinite proceeds as before, using $`𝚔_{}`$ in place of $`𝚔_0`$. If $`K`$ is finite, the proof is even closer to that for the logic without $`\mathrm{𝚂𝚎𝚕𝚏}`$. As sketched in the main text, because $`S`$ is consistent, it follows from Identity, Witnesses, and Self-is-key that there must be some key $`𝚔_{}K`$ such that $`\mathrm{𝚂𝚎𝚕𝚏}𝚔_{}S`$. For this key $`𝚔_{}`$, we must have $`𝚔_{}\text{’s}𝚗𝚔S`$ iff $`𝚗𝚔S`$. Thus, $`𝚔_{}`$ plays the role of $`𝚔_0`$ in the earlier argument. (Note that we now no longer need Current Principal to ensure the existence of $`𝚔_0`$.) The rest of the argument is unchanged.) ## Acknowledgments Work on this paper was done while the second author was with the School of Computing Sciences, University of Technology, Sydney. This work was supported in part by NSF under grant IRI-96-25901 and by a UTS internal research grant. A preliminary version of this paper appeared in the Proceedings of the 12th IEEE Computer Security Foundations Workshop, 1999, pp. 111–122.
warning/0001/cond-mat0001331.html
ar5iv
text
# Spin-triplet superconductivity in quasi-one dimension \[ ## Abstract We consider a system with electron-phonon interaction, antiferromagnetic fluctuations and disconnected open Fermi surfaces. The existence of odd-parity superconductivity in this circumstance is shown for the first time. If it is applied to the quasi-one-dimensional systems like the organic conductors $`(\mathrm{TMTSF})_2\mathrm{X}`$ we obtain spin-triplet superconductivity with nodeless gap. Our result is also valid in higher dimensions(2d and 3d). \] In recent years the signs for unconventional superconductivity in many compounds have been accumulated. Examples include high $`T_\mathrm{c}`$ superconductors, heavy fermions, organic conductors, Sr<sub>2</sub>RuO<sub>4</sub>, etc. The common features of those compounds are quasi-low dimensionality and proximity of antiferromagnetic(AF) order. Here we consider seemingly the simplest quasi-one-dimensional systems which are realized in (TMTSF)<sub>2</sub>X family where X$`=`$ PF<sub>6</sub>, AsF<sub>6</sub>, SbF<sub>6</sub>, ClO<sub>4</sub>, etc.(Bechgaard salts). At ambient pressure, most of these extremely anisotropic compounds undergo a metal-insulator transition at low temperature and have a spin-density-wave(SDW) fundamental state. Under moderate pressure, the SDW instability is suppressed and replaced by a superconducting transition at a critical temperature of order of 1K . (One exception to this is (TMTSF)<sub>2</sub>ClO<sub>4</sub> which is superconducting at ambient pressure.) Thus these compounds may be characterized by competition between superconducting and SDW ground states . As for the gap symmetry, Takigawa et al. measured nuclear relaxation rate of proton in (TMTSF)<sub>2</sub>ClO<sub>4</sub>. Their results show unconventional superconductivity(absence of coherence peak) and the existence of nodes. However their measurement is restricted to $`T>0.5T_\mathrm{c}`$ and the existence of nodes are not yet conclusive. More recently, Belin and Behnia showed some evidences for nodeless gap by thermal conductivity measurement. Also experiments on (TMTSF)<sub>2</sub>ClO<sub>4</sub> and (TMTSF)<sub>2</sub>PF<sub>6</sub> show that $`H_{\mathrm{c2}}`$ for $`\stackrel{}{B}\stackrel{}{a}`$ and $`\stackrel{}{B}\stackrel{}{b}`$ far exceed the Pauli-limiting field $`B_\mathrm{P}\mathrm{\Delta }_0/(\sqrt{2})\mu _\mathrm{B})2`$ Tesla for Bechgaard salts, where $`\mathrm{\Delta }_0`$ is the superconducting order parameter at $`T=0`$ and $`\mu _\mathrm{B}`$ is the Bohrmagneton. This suggests spin-triplet superconductivity. Maki et al. studied theoretically impurity effects and vortex states on these compounds . In this paper we consider AF magnetic coupling and electron-phonon interaction under open disconnected Fermi surfaces. The spin-triplet superconductivity is possible under electron-phonon interaction and AF magnetic coupling . We assume that AF fluctuation is not strong enough to give SDW gap. We shall show the existence of spin-triplet superconductivity in this circumstance for the first time. Our result is also applicable to higher dimensions(2d and 3d). – Open disconnected Fermi surface The Fermi surface of the system we consider is quasi-one-dimensional and consists of two separated parts, which are $`(k_\mathrm{F})_xc>0`$ and $`(k_\mathrm{F})_xc`$. ($`c`$ is a constant.) It is also supposed to be symmetric under parity transformation $`kk`$. This type of Fermi surface is realized in $`(\mathrm{TMTSF})_2\mathrm{X}`$ . The non-interacting part of the Hamiltonian of $`(\mathrm{TMTSF})_2\mathrm{X}`$ is often written $$E=2t_a\mathrm{cos}(k_xa)2t_b\mathrm{cos}(k_yb).$$ (1) Here the ratio of $`t_a`$ and $`t_b`$ is about 10:1. We take two interactions between electrons: one comes from electron-phonon coupling, the other comes from AF fluctuations. – Phonon-mediated interaction It is written $$H_{\mathrm{int}}^{\mathrm{ph}}=\underset{kk^{}q}{}\underset{\alpha \beta }{}f(q)a_{k+q,\alpha }^{}a_{k,\alpha }a_{k^{}q,\beta }^{}a_{k^{},\beta },$$ (2) where $`a`$ and $`a^{}`$ are the usual fermion operators, $`\alpha `$ and $`\beta `$ represent spin orientations and $`f(q)>0`$ has a peak at $`q=0`$. We assume $`f(q)=f(q)`$. If one considers pairing interaction between $`k`$ and $`k`$, (2) reduces to $$H_{\mathrm{int}}^{\mathrm{ph}}=\frac{1}{2}\underset{k,q}{}f(q)\left(\varphi _0^{}(k+q)\varphi _0(k)+\stackrel{}{\varphi }^{}(k+q)\stackrel{}{\varphi }(k)\right),$$ (3) where $`\varphi _0(k)=a_{k,\alpha }(\sigma _2)_{\alpha ,\beta }a_{k,\beta },`$ (4) $`\stackrel{}{\varphi }(k)=a_{k,\alpha }(\sigma _2\stackrel{}{\sigma })_{\alpha ,\beta }a_{k,\beta }.`$ (5) Here $`\varphi _0(k)`$ is the spin-singlet pairing and $`\stackrel{}{\varphi }(k)`$ is the spin-triplet one: $$a_{k,\alpha }a_{k,\beta }=\frac{1}{2}(\sigma _2)_{\alpha \beta }\varphi _0(k)+\frac{1}{2}(\stackrel{}{\sigma }\sigma _2)_{\alpha \beta }\stackrel{}{\varphi }(k).$$ (6) Since $`f(q)`$ has a peak at $`q=0`$, this interaction is approximated by $`H_{\mathrm{int}}^{\mathrm{ph}}`$ $`=`$ $`{\displaystyle \frac{1}{2}}{\displaystyle \underset{k}{}}f(0)\left(\varphi _0^{}(k)\varphi _0(k)+\stackrel{}{\varphi }^{}(k)\stackrel{}{\varphi }(k)\right)`$ (7) $`=`$ $`{\displaystyle \underset{k_x>0}{}}f(0)\left(\varphi _0^{}(k)\varphi _0(k)+\stackrel{}{\varphi }^{}(k)\stackrel{}{\varphi }(k)\right).`$ (8) Interaction due to AF fluctuations It is written $`H_{\mathrm{int}}^{\mathrm{AF}}={\displaystyle \underset{kk^{}q}{}}{\displaystyle \underset{\alpha \beta \gamma \delta }{}}J(q)\stackrel{}{\sigma }_{\alpha \beta }\stackrel{}{\sigma }_{\gamma \delta }a_{k+q,\alpha }^{}a_{k,\beta }a_{k^{}q,\gamma }^{}a_{k^{},\delta },`$ (9) where $`J(q)>0`$ represents interaction due to AF fluctuations and has a peak value at a nesting vector $`q=\pm Q`$. The nesting vector $`Q`$ connects the two separated Fermi surfaces: $`Q_x2c`$. We assume $`J(q)=J(q)`$. A similar analysis to above leads to $$H_{\mathrm{int}}^{\mathrm{AF}}=\frac{1}{2}\underset{k,q}{}J(q)\left(3\varphi _0^{}(k+q)\varphi _0(k)\stackrel{}{\varphi }^{}(k+q)\stackrel{}{\varphi }(k)\right),$$ (10) and this interaction is approximated by $`H_{\mathrm{int}}^{\mathrm{AF}}`$ $`={\displaystyle \frac{1}{2}}{\displaystyle \underset{k}{}}J(Q)\left(3\varphi _0^{}(k+Q)\varphi _0(k)\stackrel{}{\varphi }^{}(k+Q)\stackrel{}{\varphi }(k)\right)`$ (12) $`+{\displaystyle \frac{1}{2}}{\displaystyle \underset{k}{}}J(Q)\left(3\varphi _0^{}(kQ)\varphi _0(k)\stackrel{}{\varphi }^{}(kQ)\stackrel{}{\varphi }(k)\right).`$ For fermions on the Fermi surface, this becomes $`H_{\mathrm{int}}^{\mathrm{AF}}`$ (14) $`={\displaystyle \frac{1}{2}}{\displaystyle \underset{k_xc}{}}J(Q)\left(3\varphi _0^{}(k+Q)\varphi _0(k)\stackrel{}{\varphi }^{}(k+Q)\stackrel{}{\varphi }(k)\right)`$ (15) $`+{\displaystyle \frac{1}{2}}{\displaystyle \underset{k_xc}{}}J(Q)\left(3\varphi _0^{}(kQ)\varphi _0(k)\stackrel{}{\varphi }^{}(kQ)\stackrel{}{\varphi }(k)\right)`$ (16) $`={\displaystyle \underset{k_x,k_x^{}c}{}}J(Q)\left(3\varphi _0^{}(k^{})\varphi _0(k)+\stackrel{}{\varphi }^{}(k^{})\stackrel{}{\varphi }(k)\right).`$ (17) Here $`k^{}`$ satisfies $`k^{}=Qk`$. Spin-triplet superconductivity Equation(8) shows that the phonon-mediated interaction of spin-triplet pairing has the same magnitude as that of spin-singlet one. In many systems, however, only the spin-singlet superconductivity is realized. This is because the Fermi surfaces of usual matters are connected. In a system of connected Fermi surface, the requirement of parity and continuity does not admit the constant spin-triplet gap. Therefore, the s-wave superconductivity is favored from the kinematic reason. The quasi-one-dimensional system we consider, however, has disconnected Fermi surfaces, so it admits constant spin-triplet gap: $`\stackrel{}{d}|_{k_xc}=\stackrel{}{d}|_{k_xc}=\mathrm{const}.`$. <sup>*</sup><sup>*</sup>*$`\stackrel{}{d}`$ and $`\psi `$ is defined as $`\mathrm{\Delta }(k)=i\sigma _2\psi (k)+i(\stackrel{}{d(k)}\stackrel{}{\sigma })\sigma _2`$ . Therefore, there is no reason to prefer s-wave superconductivity. From (17), the interaction due to AF fluctuation disturbs the spin-singlet superconductivity more than the spin-triplet one, so the spin-triplet superconductivity is realized. This spin-triplet gap is nodeless. BCS analysis Finally, we apply the BCS weak coupling theory. Since $`T_\mathrm{c}`$ of (TMTSF)<sub>2</sub>X is rather low, one can expect that the strong coupling corrections do not change our results qualitatively. For the sake of simplicity, the cutoffs of the phonon-mediated interaction and AF one are taken to be same and denoted by $`\mathrm{}\omega _\mathrm{D}`$. We approximate the spin-singlet gap by $$\psi |_{k_xc}=\psi |_{k_xc}=\mathrm{const}.,$$ (18) and the spin-triplet gap by $$\stackrel{}{d}|_{k_xc}=\stackrel{}{d}|_{k_xc}=\mathrm{const}..$$ (19) If we denote the density of states on the Fermi surface as $`N(k_\mathrm{F})`$, the critical temperature of the spin-singlet superconductivity becomes $$T_\mathrm{c}^{\mathrm{even}}=1.13\mathrm{}\omega _\mathrm{D}e^{1/\{N(k_\mathrm{F})V^{\mathrm{even}}\}},$$ (20) where $$V^{\mathrm{even}}=f(0)+3J(Q),$$ (21) and that of spin-triplet becomes $$T_\mathrm{c}^{\mathrm{odd}}=1.13\mathrm{}\omega _\mathrm{D}e^{1/\{N(k_\mathrm{F})V^{\mathrm{odd}}\}},$$ (22) where $$V^{\mathrm{odd}}=f(0)+J(Q).$$ (23) From (21) and (23), the condition $`V^{\mathrm{odd}}<V^{\mathrm{even}}`$ is always satisfied. Thus $`T_\mathrm{c}^{\mathrm{odd}}>T_\mathrm{c}^{\mathrm{even}}`$ and spin-triplet superconductivity is present if $`V^{\mathrm{odd}}=f(0)+J(Q)<0`$. This simple result show that there is no superconductivity if AF fluctuations dominate and we have spin-triplet superconductivity if phonon-mediated interaction dominates. Since $`\stackrel{}{d}(k)`$ is an odd function, the gaps at $`(k_\mathrm{F})_x=\pm c`$ have opposite signs and there is no node in the gap.
warning/0001/astro-ph0001366.html
ar5iv
text
# The abundance of high-redshift objects as a probe of non–Gaussian initial conditions ## 1 Introduction In the standard inflationary model for triggering structure formation in the Universe, there are precise predictions for the properties of the initial fluctuations: they are adiabatic and follow a nearly-Gaussian distribution, with deviations from Gaussianity which are calculable, small and generally dependent on the specific inflationary model (Falk et al., 1993; Gangui et al., 1994; Gangui, 1994; Wang & Kamionkowski, 1999; Gangui & Martin, 1999). Because of the smallness of such deviations from the Gaussian behaviour, in most theoretical predictions one simply assumes that the primordial density field has exactly random phases. Consequently, intrinsic temperature fluctuations in the Cosmic Microwave Background (CMB) are commonly treated as being Gaussian and the same assumption is made in most analyses of Large-Scale Structure (LSS) of the Universe. Besides this ‘standard’ nearly-Gaussian model for generating cosmological structures, based on the amplification of quantum fluctuations of the same scalar field which drives the inflationary dynamics, there exist alternative models for the origin of fluctuations which predict stronger deviations from the random-phase paradigm. Still within the context of inflation multiple scalar field models can give rise to non-Gaussian perturbations of either isocurvature or adiabatic type (Allen et al., 1987; Kofman & Pogosyan, 1988; Salopek et al., 1989; Linde & Mukhanov, 1997; Peebles, 1999a, b; Salopek, 1999). Alternatively, cosmological defect scenarios (e.g. Vilenkin 1985; Vachaspati 1986; Hill et al. 1989; Turok 1989; Albrecht & Stebbins 1992) generally predict non-Gaussian initial conditions. The observed abundance of high-redshift cosmic structures contains important information about the properties of initial conditions on galaxy and clusters scales. CMB observations will put constraints on the nature of the initial conditions (e.g. Pen & Spergel 1995; Hu et al. 1997; Verde et al. 1999; Wang & Kamionkowski 1999; Wang et al. 1999), but will be severely affected by the presence of noise and foregrounds (e.g. Knox 1999; Tegmark 1998; Bouchet & Gispart 1999) and probe rather large scales. The Gaussian assumption plays a central role in analytical predictions for the abundance and statistical properties of the first objects to collapse in the Universe. In this context, the formalism proposed by Press & Schechter (Press & Schechter, 1974), with its later extensions and improvements (Peacock & Heavens, 1990; Bond et al., 1991; Cole, 1991; Lacey & Cole, 1993) has become the ‘standard lore’ for predicting the number of collapsed dark matter halos as a function of redshift. However, even a small deviation from Gaussianity would have a deep impact on those statistics which probe the tails of the distribution. This is indeed the case for the abundance of high-redshift objects like galaxies at $`z\begin{array}{c}>\hfill \\ \hfill \end{array}5`$ or clusters at $`z1`$ which correspond to high peaks, i.e. rare events, in the underlying dark matter density field. Therefore, even small deviations from Gaussianity might be potentially detectable by looking at the statistics of high-redshift systems. The importance of using the mass-function as a tool to distinguish among different non-Gaussian statistics for the primordial density field, was first recognized by Lucchin & Matarrese (1988), Colafrancesco et al. (1989) and, more recently, by Chiu et al. (1998), followed by Robinson & Baker (1999), Robinson et al. (1999a, b), Koyama et al. (1999), Willick (1999), Avelino & Viana (1999). To make predictions on the number counts of high-redshift structures in the context of non-Gaussian initial conditions, a generalized version of the PS approach had to be introduced. Such a generalization of the PS formalism has been tested successfully against N-body simulations (Robinson & Baker, 1999), but, from a theoretical point of view, it suffers from the same problems of the original Press & Schechter (PS) formulation: it cannot properly account for the so called cloud-in-cloud problem, i.e. the constraint that bound systems of given mass should not be incorporated in larger mass condensations of the same catalog<sup>1</sup><sup>1</sup>1The so called cloud-in-cloud problem arises only if one is interested in predicting the local density of maxima; this feature is not a problem if one focuses the attention to percolation regions where the ratio of the local density $`\rho `$ to the background density $`\rho _b`$ is $`\rho /\rho _b200`$.. Even more important, most of the PS generalizations to non-Gaussian models proposed in the literature do not properly take into account the dependence of the fluctuation field on the smoothing scale. In this paper we obtain an analytic prediction for the number of dark matter halos as a function of redshift, within the hierarchical structure formation paradigm, following the PS formalism. The strength of our method is that we are able to properly take into account the smoothing-scale, or mass, dependence of the probability distribution function of the primordial density contrast. Obtaining analytical results in this context is extremely important. Direct simulations of non-Gaussian fields are generally plagued by the difficulty of properly accounting for the non-linear way in which resolution and finite box-size effects, present in any realization of the underlying Gaussian process, propagate into the statistical properties of the non-Gaussian field. Moreover, finite volume realizations of non-Gaussian fields might fail in producing fair samples of the assumed statistical distribution, i.e. ensemble and (finite-volume) spatial distributions might sensibly differ (e.g. Zel’dovich et al. 1987). This problem, of course, becomes exacerbated and hard to keep under control in so far as the tails of the distribution are concerned. Thus, in looking for the likelihood of rare events for a non-Gaussian density field, either exact or approximate analytical estimates should be considered as the primary tool. When compared with N-body simulation results the striking feature of the standard PS algorithm is that it works extremely well in predicting properties of highly non-linear objects such as the DM haloes, even being based on linear theory and relatively simple assumptions. This agreement has been proven for Gaussian initial conditions, and one might wonder whether such a feature also extends to the non-Gaussian case, where additional non-linear couplings arise and non-linear gravitational corrections can be important at larger scales, unlike the Gaussian case. This extension of the PS prediction will be tested against N-body simulations in a subsequent paper. For the present application, however, the primordial non-Gaussianity is assumed to be small and we can safely assume that departures from the PS prediction are negligible. The outline of the paper is as follows: in Section 2 we propose a parameterization of primordial non-Gaussianity that covers a wide range of physically motivated models whose non-Gaussianity can be dialed from zero (the Gaussian limit). We will then relate analytically the non-Gaussianity parameter to the number of high-redshift objects. This step involves the generalization of the PS formalism to non-Gaussian initial conditions, which in turn requires an expression for the probability density function of the non-Gaussian mass-density field as a function of the filtering radius. In Section 3, we calculate the probability density function for the density fluctuation field, smoothed on the mass scale typical of high-redshift objects. Since the non-Gaussianity is expected to be small, we can calculate the probability density function by expanding the ‘cumulant generator’ in powers of the non-Gaussianity parameter and keeping only linear order terms. In this context we are able to obtain an analytic expression for the probability density function of the smoothed density field. We generalize the PS approach to non-Gaussian density fields in Section 4, where we give an analytical expression for the comoving mass-function of halos formed at a given redshift, also accounting for the cloud-in-cloud problem. Section 5 discusses the association between halos and high-redshift galaxies. The PS formalism predicts the number of dark matter halos, but the correspondence to the observed number of galaxies is not necessarily simple. Finally in Section 6 we summarize our main results and draw our conclusions. ## 2 Parameterization of primordial non-Gaussianity Because of the infinite range of possible non-Gaussian models, we consider here two models with primordial non-Gaussianity, whose amplitude can be dialed from zero (the Gaussian limit). Following Verde et al. (1999) we consider models in which either the density contrast (model A) or the gravitational potential (model B) contains a part which is the square of a Gaussian random field. The physical motivation for this choice is that such a non-Gaussianity may arise in slow-roll and/or nonstandard (e.g. two-field) inflation models (Luo, 1994; Falk et al., 1993; Gangui et al., 1994; Fan & Bardeen, 1992); moreover both models may be considered as a Taylor expansion of more general non-Gaussian fields (e.g. Coles & Barrow 1987; Verde et al. 1999), and are thus fairly generic forms of mild non-Gaussianity. We note here that, since in the PS framework the evolution of perturbations is considered to be linear, this parameterization of non-Gaussianity is effectively equivalent to non-linear biasing acting on a truly Gaussian underlying field. The effect of biasing is to alter the clustering properties of the galaxies with respect to the dark matter by relating galaxy formation efficiency to the environment. We do not intend to investigate the clustering properties of high-redshift objects, we will instead focus on the probability density function (PDF) of the linear density contrast as a tool to predict the abundance of dark matter halos as a function of mass and formation redshift. ### 2.1 Linear plus quadratic model for the density or the gravitational potential As a first step, we want to obtain the PDF for a model where either the primordial over-density field $`\delta \delta \rho /\rho `$, or the initial peculiar gravitational potential $`\mathrm{\Phi }`$ is represented by a zero-mean random field $`\psi `$, given by the following local transformation $``$ on an underlying Gaussian field $`\varphi `$: $$\psi (𝐱)=[\varphi ]\alpha \varphi (𝐱)+ϵ(\varphi ^2(𝐱)\varphi ^2),$$ (1) with $`\alpha `$ and $`ϵ`$ free parameters of the model. The homogeneous and isotropic Gaussian process $`\varphi `$ is assumed to have zero mean, $`\varphi =0`$, and power-spectrum $`P_\varphi `$ to be specified later. Note that in the limit $`\alpha 0`$, $`\psi `$ is chi-squared distributed, while for $`ϵ0`$ one recovers the Gaussian case. As pointed out before, the model of eq. (1) can be thought as representing the first two terms of the Taylor expansion of more general non-Gaussian fields around the Gaussian limit. Cubic or higher order powers of $`\varphi `$ in such an expansion should be thought as being of order $`ϵ^2`$ or higher. We will come back to this point in Section 3.3. Since the overall normalization of $`\psi `$ is fixed observationally (e.g. by measurements of the power-spectrum), effectively there is only one free parameter in the model. In what follows we will take either $`\alpha =1`$ and $`ϵ`$ as the non-Gaussianity parameter or $`\alpha =0`$ and $`ϵ=1`$ to obtain the $`\chi ^2`$ model. If there was no filtering to take into account or if the transformation (1) were true for a smoothed field, then the PDF for $`\psi `$ could be simply obtained as follows $$P(\psi )\delta ^D(\alpha \varphi +ϵ(\varphi ^2\varphi ^2)\psi )𝑑\varphi P(\varphi )\delta ^D(\alpha \varphi +ϵ(\varphi ^2\varphi ^2)\psi )$$ (2) where $`P(\varphi )`$ is a Gaussian PDF with zero mean and variance $`\varphi ^2`$ and $`\delta ^D`$ is the Dirac delta function. The above equation can be also thought as resulting from applying the Chapman-Kolmogorov equation: $`P(\psi )=𝑑\varphi W(\psi |\varphi )P(\varphi )`$ where $`W(\psi |\varphi )=\delta ^D(\psi [\varphi ])`$ is the transition probability from $`\varphi `$ to $`\psi `$ (e.g. Taylor & Watts 2000). The integral over $`\varphi `$ can be performed by writing: $$P(\psi )=𝑑\varphi P(\varphi )\delta ^D(\varphi ^1[\psi ])\frac{1}{d/d\varphi },$$ (3) which gives: $$P(\psi )=\left\{2\pi \varphi ^2\left[\alpha ^2+4ϵ\left(ϵ\varphi ^2+\psi \right)\right]\right\}^{1/2}\left[_{}(\psi )_+(\psi )\right],$$ (4) with $$_\pm (\psi )\mathrm{exp}\left\{\frac{1}{ϵ^2\varphi ^2}\left[\alpha ^2+2ϵ\left(ϵ\varphi ^2+\psi \right)\pm \alpha \sqrt{\alpha ^2+4ϵ(ϵ\varphi ^2+\psi )}\right]\right\}.$$ (5) However, in order to apply the PS formalism, we need to know the PDF for the linear mass-density field as a function of the smoothing radius $`R`$ and redshift $`z`$. To take into account the different evolution of modes inside the horizon, we assume that the transfer function acts on our primordial non-Gaussian field $`\psi (𝐱)`$, as a convolution. We can easily account for the smoothing operation, the effect of the transfer function and the linear growth of perturbations, by writing (e.g. Moscardini et al. 1991): $$\delta _R(𝐱,z)=D(z)\delta _R(𝐱)=D(z)d^3yF_R(|𝐱𝐲|)\psi (𝐲),$$ (6) where $`D(z)`$ is the growing mode of linear perturbations, normalized to unity at $`z=0`$, so that $`\delta _R(𝐱)`$ represents the mass density fluctuation field linearly extrapolated to the present time. The isotropic function $`F_R(|𝐱|)`$ can be specified by its Fourier transform, $$\stackrel{~}{F}_R(k)d^3ye^{i𝐤𝐲}F(|𝐲|)=\stackrel{~}{W}_R(k)T(k)g(k),$$ (7) where $`\stackrel{~}{W}_R(k)`$ is the Fourier transform of the assumed low-pass filter (e.g. a spherical top-hat filter), $`T(k)`$ is the transfer function (normalized to unity for $`k0`$), that we will later assume to take the adiabatic CDM form as in (Bardeen et al., 1986), with the modifications of Sugiyama (1995). The function $`g(k)`$ completes the specification of our model: * model A: $`g(k)=1`$, if we assume that our non-Gaussian model applies directly to the primordial density field. * model B: $`g(k)=\frac{2}{3}(k/H_0)^2\mathrm{\Omega }_{0m}^1`$, if the same assumption is made for the gravitational potential (its precise form comes from solving the cosmic Poisson equation); here $`\mathrm{\Omega }_{0m}`$ denotes the present day density parameter of non-relativistic – baryonic plus dark – matter. As widely discussed by Moscardini et al. (1991), the overall sign of the Gaussian to non-Gaussian mapping inherent in these models is a crucial parameter, which determines most of the non-linear dynamics. In our case, we are still free to choose the sign, through the actual choice of the free parameter $`ϵ`$. We finally get $$\delta _R(𝐱)=\alpha \varphi _R(𝐱)+ϵd^3yF_R(𝐱𝐲)\varphi ^2(𝐲)C,$$ (8) where $`\varphi _R(𝐱)`$ is just the smoothed underlying Gaussian field i.e. the convolution of $`\varphi `$ with $`F`$, and $$Cϵ\varphi ^2d^3yF_R(𝐱𝐲)$$ (9) ensures that $`\delta _R`$ has zero expectation value. ## 3 PDF of the smoothed density fluctuations In this section we derive an approximate analytic expression, valid for small values of the non-Gaussian parameter $`ϵ`$, for the PDF as a function of the smoothing radius (i.e. mass scale) and of the redshift of collapse. This is the key ingredient to develop the extension of the PS formalism to non-Gaussian fields. The reader mainly interested in the application of the PS approach to non-Gaussian density fields may omit reading this section at a first sitting. Once the filtering process has been taken into account, the density fluctuation PDF can be obtained through a functional or path-integral <sup>2</sup><sup>2</sup>2The path-integral approach has been widely applied in the cosmological context and in particular to large-scale structure studies by e.g. Politzer & Wise (1984); Grinstein & Wise (1986); Matarrese, Lucchin & Bonometto (1986); Bertschinger (1987). (e.g. Ramond 1989) $`P(\delta _R)`$ $`=`$ $`\delta ^D\left(\alpha \varphi _R(𝐱)+ϵ{\displaystyle d^3yF_R(𝐱𝐲)\varphi ^2(𝐲)}C\delta _R(𝐱)\right)`$ $`=`$ $`{\displaystyle [𝒟\varphi ]𝒫[\varphi ]\frac{d\lambda }{2\pi }\mathrm{exp}\left[i\lambda \left(\alpha \varphi _R(𝐱)+ϵd^3yF_R(𝐱𝐲)\varphi ^2(𝐲)C\delta _R(𝐱)\right)\right]},`$ where, in the second line, we used the integral representation of the Dirac delta function. The functional integration is over $`\varphi `$ configurations in real space weighted by the Gaussian probability density functional $$𝒫[\varphi ]=\mathrm{exp}\left\{\frac{1}{2}d^3yd^3z\varphi (𝐲)𝒦(𝐲,𝐳)\varphi (𝐳)\right\}/[𝒟\varphi ]\mathrm{exp}\left\{\frac{1}{2}d^3yd^3z\varphi (𝐲)𝒦(𝐲,𝐳)\varphi (𝐳)\right\},$$ (11) which has been consistently normalized to unit total probability, $`[𝒟\varphi ]𝒫[\varphi ]=1`$. Although the previous expressions for the PDF actually gives its form at redshift $`z=0`$, one should keep in mind that the quantity $`P(\delta _R)d\delta _R`$ is redshift-independent, as long as linear evolution applies. From now on we will use the following compact notation: $$d^3yd^3z\varphi (𝐲)𝒦(𝐲,𝐳)\varphi (𝐳)(\varphi ,𝒦,\varphi )$$ (12) The symmetric kernel $`𝒦`$ is defined as the functional inverse of the two-point correlation function $`\xi _\varphi (𝐲,𝐰)=\xi _\varphi (|𝐲𝐰|)`$ of the field $`\varphi `$, namely, $$d^3y𝒦(𝐳,𝐲)\xi _\varphi (𝐲,𝐰)=\delta ^D(𝐳𝐰).$$ (13) By exploiting the fact that $$d^3yF_R(𝐱𝐲)\varphi ^2(𝐲)=d^3yd^3z\varphi (𝐲)F_R(𝐱𝐲)\delta ^D(𝐲𝐳)\varphi (𝐳)(\varphi ,F_R\delta ^D,\varphi ),$$ (14) equation (3) becomes: $$P(\delta _R)=\frac{\frac{d\lambda }{2\pi }e^{i\lambda \delta _Ri\lambda C}[𝒟\varphi ]\mathrm{exp}\left\{\frac{1}{2}(\varphi ,𝒦,\varphi )+i\lambda ϵ(\varphi ,F_R\delta ^D,\varphi )+i(𝒥_\lambda ,\varphi )\right\}}{[𝒟\varphi ]e^{\frac{1}{2}(\varphi ,𝒦,\varphi )}},$$ (15) where we have defined the source functional $$𝒥_\lambda (𝐲)\lambda \alpha F_R(𝐱𝐲)$$ (16) and introduced the notation $$(𝒥_\lambda ,\varphi )d^3y𝒥_\lambda (𝐲)\varphi (𝐲).$$ (17) The above functional integration can be performed analytically, by applying the so-called path-integral technique for composite operators (Cornwall et al., 1974; Hawking & Moss, 1983). Let us briefly sketch the main steps of the procedure. We start by defining a new kernel: $$𝒦_\lambda ^{}(𝐳,𝐲)𝒦(𝐳,𝐲)i2\lambda ϵF_R(𝐲𝐱)\delta ^D(𝐳𝐲),$$ (18) which allows to write $$P(\delta _R)=\frac{d\lambda }{2\pi }e^{i\lambda \delta _Ri\lambda C}[𝒟\varphi ]e^{\frac{1}{2}(\varphi ,𝒦_\lambda ^{},\varphi )+i(𝒥_\lambda ,\varphi )}.$$ (19) We then make the change of variable (under which the path-integral is left unchanged) $$\varphi (𝐳)\varphi (𝐳)id^3y\left[𝒦_\lambda ^{}\right]^1(𝐳,𝐲)𝒥_\lambda (𝐲)$$ (20) where $`\left[𝒦_\lambda ^{}\right]^1`$ satisfies the integral equation $$d^3y𝒦^{}(𝐳,𝐲)\left[𝒦_\lambda ^{}\right]^1(𝐲,𝐰)=\delta ^D(𝐳𝐰).$$ (21) We then easily get: $$P(\delta _R)=\frac{d\lambda }{2\pi }e^{i\lambda \delta _Ri\lambda C\frac{1}{2}(𝒥_\lambda ,\left[𝒦_\lambda ^{}\right]^1,𝒥_\lambda )}\left\{\mathrm{Det}[𝒦_\lambda ^{}]/\mathrm{Det}[𝒦]\right\}^{1/2},$$ (22) having used the standard notation for functional determinants: $$\{Det[𝒦_\lambda ^{}]/Det[𝒦]\}^{1/2}\frac{[𝒟\varphi ]\mathrm{exp}[\frac{1}{2}(\varphi ,𝒦_\lambda ^{},\varphi )]}{[𝒟\varphi ]\mathrm{exp}[\frac{1}{2}(\varphi ,𝒦,\varphi )}=\mathrm{exp}\left\{\frac{1}{2}\mathrm{Tr}\mathrm{ln}\left[\mathrm{𝟏}i2\lambda ϵ(F_R\delta ^D,𝒦^1)\right]\right\}.$$ (23) The reader unfamiliar with the functional notation can understand the last result as a generalization of the identity $`\mathrm{ln}[\mathrm{det}𝐌]=\mathrm{Tr}[\mathrm{ln}𝐌]`$, which applies to any symmetric matrix $`𝐌`$. Here $`\mathrm{𝟏}`$ denotes the functional unit matrix, i.e. the Dirac delta function; the logarithm of a functional is defined as its series expansion and $$(F_R\delta ^D,𝒦^1)d^3wF_R(|𝐰𝐱|)\delta ^D(𝐲𝐰)𝒦^1(𝐰𝐳)=F_R(|𝐲𝐱|)\xi _\varphi (𝐲𝐳).$$ (24) Finally, the trace (Tr) of a functional $`G(𝐲,𝐳)`$ is defined as $`d^3yd^3z\delta ^D(𝐲𝐳)G(𝐲,𝐳)`$. Therefore, equation (23) becomes: $$\{Det[𝒦_\lambda ^{}]/Det[𝒦]\}^{1/2}=\mathrm{exp}\left\{\frac{1}{2}d^3yd^3z\delta ^D(𝐲𝐳)\mathrm{ln}\left[\delta ^D(𝐲𝐳)i2\lambda ϵF_R(|𝐲𝐱|)\xi _\varphi (𝐲𝐳)\right]\right\}$$ (25) The PDF for the linearly evolved, smoothed density field takes the exact form: $$P(\delta _R)d\delta _R=\frac{d\lambda }{2\pi }e^{i\lambda \delta _R+𝒲(\lambda )}𝑑\delta _R,$$ (26) where $`𝒲(\lambda )`$ is called the cumulant generator, as its series expansion around $`\lambda =0`$ defines the cumulants (or irreducible moments) of $`\delta _R(𝐱)`$. Its exact form is: $`𝒲(\lambda )`$ $`=`$ $`i\lambda C{\displaystyle \frac{1}{2}}{\displaystyle }d^3y{\displaystyle }d^3z\{\lambda ^2\alpha ^2F_R(|𝐲|)\left[𝒦_\lambda ^{}\right]^1(𝐲,𝐳)F_R(|𝐳|)`$ (27) $`+`$ $`\delta ^D(𝐲𝐳)\mathrm{ln}[\delta ^D(𝐲𝐳)i2\lambda ϵF_R(|𝐲|)\xi _\varphi (𝐲𝐳)]\}.`$ where the $`\lambda `$ subscript indicates where the $`\lambda `$ dependence is hidden. Note that the $`𝐱`$ dependence has been eliminated by a mere translation of the origin. Equation (27) contains also the exact form for the generating function for the particular case where the original non-Gaussian field $`\psi `$ is chi-squared distributed: this is simply obtained by setting $`\alpha =0`$ in the previous expression. Note also that the first-order term in the expansion of the logarithm precisely cancels the $`i\lambda C`$ term, so that the condition $`\delta _R=0`$ is identically satisfied. ### 3.1 Kernel inversion In order to solve the remaining integrals we need to find an expression for the functional $`[𝒦_\lambda ^{}]^1`$. Let us start by finding an expression for the kernel $`𝒦`$, given that its inverse is just the auto-correlation function of the Gaussian field $`\varphi `$, namely, $$d^3y𝒦(𝐰𝐲)\xi _\varphi (𝐲𝐳)=\delta ^D(𝐰𝐳).$$ (28) Fourier transforming both sides of eq. (28) we easily find (Politzer & Wise, 1984; Bertschinger, 1987) $$\stackrel{~}{𝒦}(k)=1/P_\varphi (k)$$ (29) where $`P_\varphi (k)`$ is the power-spectrum of the underlying Gaussian field $`\varphi `$. Therefore $`𝒦(𝐰,𝐲)𝒦(𝐰𝐲)𝒦(r)`$ reads $$𝒦(r)=\frac{1}{(2\pi )^3}d^3k\mathrm{exp}(i𝐤𝐫)\frac{1}{P_\varphi (k)}=\frac{1}{2\pi ^2}_0^{\mathrm{}}𝑑kk^2j_0(kr)\frac{1}{P_\varphi (k)}.$$ (30) To find an expression for $`\left[𝒦_\lambda ^{}\right]^1`$ one would like to proceed in an analogous way. However, owing to the absence of translational invariance of $`𝒦^{}`$, going to momentum space does not help. The technique we are then going to use is to expand $`\left[𝒦_\lambda ^{}\right]^1`$ in powers of the non-Gaussianity parameter $`ϵ`$, as follows: $$\left[𝒦_\lambda ^{}\right]^1(𝐲,𝐳)=\underset{n=0}{\overset{\mathrm{}}{}}(2iϵ\lambda )^n^{(n)}(𝐲,𝐳)$$ (31) where the coefficients $`^{(n)}`$ are obtained recursively from $$^{(n)}(𝐲,𝐳)=d^3w\xi _\varphi (|𝐲𝐰|)F_R(|𝐰|)^{(n1)}(𝐰,𝐳)(n>0),$$ (32) with $`^{(0)}(𝐲,𝐳)=\xi _\varphi (|𝐲𝐳|)`$. Similarly, in Fourier space we have $$\stackrel{~}{}^{(n)}(𝐤,𝐤^{})=P_\varphi (k)\frac{d^3q}{(2\pi )^3}\stackrel{~}{F}_R(q)\stackrel{~}{}^{(n1)}(𝐪+𝐤,𝐤^{}),$$ (33) with $`\stackrel{~}{}^{(0)}(𝐤,𝐤^{})=(2\pi )^3P_\varphi (k)\delta ^D(𝐤+𝐤^{})`$. From equation (33) we find: $$\stackrel{~}{}^{(1)}(𝐤,𝐤^{})=P_\varphi (k)P_\varphi (k^{})\stackrel{~}{F}_R(|𝐤+𝐤^{}|)$$ (34) and, for $`n2`$, $`\stackrel{~}{}^{(n)}(𝐤,𝐤^{})`$ $`=`$ $`P_\varphi (k)P_\varphi (k^{}){\displaystyle \frac{d^3q_1}{(2\pi )^3}\mathrm{}\frac{d^3q_{n1}}{(2\pi )^3}\stackrel{~}{F}_R(q_1)\mathrm{}\stackrel{~}{F}_R(q_{n1})}`$ $`\times `$ $`\stackrel{~}{F}_R(|𝐪_1+\mathrm{}+𝐪_{n1}+𝐤+𝐤^{}|)P_\varphi (|𝐪_1+𝐤|)\mathrm{}P_\varphi (|𝐪_1+\mathrm{}+𝐪_{n1}+𝐤|).`$ Note that, while in principle the expressions for $`𝒦^{}`$, $`^{(n)}`$ and $`\stackrel{~}{}^{(n)}`$ should be symmetrized, for the purpose of calculating the PDF or the cumulants, this operation is not needed. ### 3.2 Cumulant generator At this point we are ready to provide an expression for the cumulant generator, by first expanding it in powers of $`\lambda `$. We have: $$𝒲(\lambda )\underset{n=2}{\overset{\mathrm{}}{}}\frac{(i\lambda )^n}{n!}\mu _{n,R},$$ (36) where $`\mu _{n,R}`$ denotes the cumulant of order $`n`$ of the smoothed density contrast $`\delta _R`$. The variance, the skewness and the kurtosis of the smoothed non-Gaussian density field are, respectively <sup>3</sup><sup>3</sup>3The integral over k in the sub-leading term of the variance can diverge for certain choices of the power-spectrum of the underlying Gaussian field and transfer function. In these cases one should bear in mind that any physical process originating the underlying field will necessarily provide the $`\varphi `$ power-spectrum with both infrared and ultraviolet cutoffs (the present-day horizon and the reheating scale, respectively, in the case of inflation-generated perturbations). In case this contribution dominates over the leading-order contribution even for $`ϵ1`$ one can postulate the existence of a term proportional to $`ϵ^2`$ in eq. (1) that will cancel out the sub-leading term in the variance. Alternatively eq. (1) and eq. (37) can be renormalized by choosing $`\alpha `$ and $`A_\varphi `$ such that $`\mu _{2,R}\mu _{2,R}^{(1)}`$ when $`ϵ1`$. Our calculations still apply in this case provided one interprets $`ϵ`$ as $`ϵ/\alpha ^2`$., $`\mu _{2,R}`$ $``$ $`\sigma _R^2\delta _R^2={\displaystyle \frac{\alpha ^2}{2\pi ^2}}{\displaystyle _0^{\mathrm{}}}𝑑kk^2\stackrel{~}{F}_R^2(k)P_\varphi (k)`$ (37) $`+`$ $`{\displaystyle \frac{ϵ^2}{2\pi ^4}}{\displaystyle _0^{\mathrm{}}}𝑑kk^2P_\varphi (k){\displaystyle _0^{\mathrm{}}}𝑑k^{}k^^2P_\varphi (k^{}){\displaystyle _0^1}𝑑\mu \stackrel{~}{F}_R^2(\sqrt{k^2+k^^2+2kk^{}\mu }),`$ $`\mu _{3,R}`$ $``$ $`\delta _R^3={\displaystyle \frac{3ϵ\alpha ^2}{2\pi ^4}}{\displaystyle _0^{\mathrm{}}}𝑑kk^2\stackrel{~}{F}_R(k)P_\varphi (k){\displaystyle _0^{\mathrm{}}}𝑑k^{}k^^2\stackrel{~}{F}_R(k^{})P_\varphi (k^{}){\displaystyle _0^1}𝑑\mu \stackrel{~}{F}_R(\sqrt{k^2+k^^2+2kk^{}\mu })`$ $`+`$ $`8ϵ^3{\displaystyle \frac{d^3k_1}{(2\pi )^3}\frac{d^3k_2}{(2\pi )^3}\frac{d^3k_3}{(2\pi )^3}P_\varphi (k_1)P_\varphi (k_2)P_\varphi (k_3)\stackrel{~}{F}_R(|𝐤_1𝐤_2|)\stackrel{~}{F}_R(|𝐤_2𝐤_3|)\stackrel{~}{F}_R(|𝐤_3𝐤_1|)}.`$ and $`\mu _{4,R}`$ $``$ $`\delta _R^43\delta _R^2^2`$ $`=`$ $`48ϵ^2\alpha ^2{\displaystyle \frac{d^3k_1}{(2\pi )^3}\frac{d^3k_2}{(2\pi )^3}\frac{d^3k_3}{(2\pi )^3}P_\varphi (k_1)P_\varphi (k_2)P_\varphi (|𝐤_3+𝐤_1|)\stackrel{~}{F}_R(k_1)\stackrel{~}{F}_R(k_2)\stackrel{~}{F}_R(k_3)}`$ $`\times `$ $`\stackrel{~}{F}_R(|𝐤_1+𝐤_2+𝐤_3|)+96ϵ^4{\displaystyle \frac{d^3k_1}{(2\pi )^3}\frac{d^3k_2}{(2\pi )^3}\frac{d^3k_3}{(2\pi )^3}\frac{d^3k_4}{(2\pi )^3}P_\varphi (k_1)P_\varphi (k_2)P_\varphi (k_3)P_\varphi (k_4)}`$ $`\times `$ $`\stackrel{~}{F}_R(|𝐤_1𝐤_2|)\stackrel{~}{F}_R(|𝐤_2𝐤_3|)\stackrel{~}{F}_R(|𝐤_3𝐤_4|)\stackrel{~}{F}_R(|𝐤_4𝐤_1|).`$ More in general, the cumulants are made of two contributions: a leading term of order $`ϵ^{n2}`$ plus a subleading term of order $`ϵ^n`$, $$\mu _{n,R}ϵ^{n2}\mu _{n,R}^{(1)}+ϵ^n\mu _{n,R}^{(2)},$$ (40) where $$\mu _{n,R}^{(1)}=2^{n3}n!\alpha ^2\frac{d^3k}{(2\pi )^3}\frac{d^3k^{}}{(2\pi )^3}\stackrel{~}{F}_R(k_1)\stackrel{~}{F}_R(k_2)\stackrel{~}{}^{(n2)}(𝐤,𝐤^{})$$ (41) and $$\mu _{n,R}^{(2)}=2^{n1}(n1)!\frac{d^3k_1}{(2\pi )^3}\mathrm{}\frac{d^3k_n}{(2\pi )^3}P_\varphi (k_1)\mathrm{}P_\varphi (k_n)\stackrel{~}{F}_R(|𝐤_1𝐤_2|)\mathrm{}\stackrel{~}{F}_R(|𝐤_{n1}𝐤_n|)\stackrel{~}{F}_R(|𝐤_n𝐤_1|).$$ (42) Note that, as anticipated, the sign of the parameter $`ϵ`$, for a given model, fully determines the sign of the skewness as well as of all the other odd-order cumulants. The cumulants for the filtered chi-squared model are immediately recovered by taking $`\alpha =0`$ and $`ϵ=1`$ in the previous expressions, so that $`\mu _{n,R}=\mu _{n,R}^{(2)}`$. It is worth noticing that eq. (42) supplies all the cumulants of a chi-squared field smoothed on scale $`R`$. This is an interesting result for models such as that recently proposed by Peebles (1999a, b), where non-Gaussian isocurvature fluctuations are obtained with a chi-squared distributed density field. #### 3.2.1 Linear order in $`ϵ`$ To first order, the skewness depends linearly on $`ϵ`$, therefore let us define $$S_{3,R}ϵS_{3,R}^{(1)}=ϵ\mu _{3,R}^{(1)}/(\mu _{2,R}^{(1)})^2,$$ (43) where $$\mu _{2,R}^{(1)}=\frac{\alpha ^2}{2\pi ^2}_0^{\mathrm{}}𝑑kk^2\stackrel{~}{F}_R^2(k)P_\varphi (k)$$ (44) and $$\mu _{3,R}^{(1)}=\delta _R^3=\frac{3\alpha ^2}{2\pi ^4}_0^{\mathrm{}}𝑑kk^2\stackrel{~}{F}_R(k)P_\varphi (k)_0^{\mathrm{}}𝑑k^{}k^^2\stackrel{~}{F}_R(k^{})P_\varphi (k^{})_0^1𝑑\mu \stackrel{~}{F}_R(\sqrt{k^2+k^^2+2kk^{}\mu })$$ (45) To this order in $`ϵ`$ there is no contribution from higher order moments. On very large scales, where the transfer function becomes unity, and for a power-law $`\varphi `$ power-spectrum ($`P_\varphi =A_\varphi k^n`$ or $`P_\varphi =A_\varphi k^{n3}`$ for model A or B, respectively), this leading-order skewness parameter $`S_{3,R}`$ becomes scale-independent, thus mimicking the behaviour induced by the gravitational instability (Peebles, 1980). For model A the variance, the skewness $`\mu _{3,R}^{(1)}`$ and skewness parameter $`S_{3,R}`$ are plotted vs. the radius $`R`$ and the mass $`M`$ in solar masses, $`MR^3`$, in Figure 1 with the following assumptions: $`\alpha =1`$, top-hat filter in real space, transfer function as in (Sugiyama, 1995), with baryonic matter density parameter $`\mathrm{\Omega }_{0b}=0.015h^2`$, $`h=0.65`$, $`\mathrm{\Omega }_{0m}=0.3`$, and scale-invariant primordial power-spectrum $`P_\varphi k`$, normalized so that the present day r.m.s. fluctuation on a sphere of $`8h^1`$ Mpc is $`\sigma _8=0.99`$ (Viana & Liddle, 1998), which simultaneously allows to best fit the local cluster abundance and the COBE data (e.g. Tegmark 1996). In what follows we will always assume that the approximation $`\sigma _R^2\mu _{2,R}^{(1)}`$ applies on all considered scales. As detailed below, at $`z=0`$, for $`ϵ0.01`$ in model A we will recover the standard PS mass-function on clusters scales, so that the relation between the observed abundance of clusters and the value of $`\sigma _8`$ keeps unchanged, in spite of our non-Gaussian assumption. To deal with model B we take $`P_\varphi k^3`$ and, as above, we normalize the mass-density variance as in model A, with the same choice of cosmological parameters and transfer function. The variance, the skewness $`\mu _{3,R}^{(1)}`$ and skewness parameter $`S_{3,R}^{(1)}`$ are plotted vs. the radius $`R`$ and the mass $`M`$ in solar masses, $`MR^3`$, in Figure 2 with the same assumptions as for model A in Figure 1. The variance for model B to linear order in $`ϵ`$ is identical to the variance for model A, but the skewness and skewness parameter are different: in particular notice that $`S_{3,R}^{(1)}`$ for model B has the opposite sign of $`ϵ`$ and, for a given value of $`ϵ`$, its amplitude is many orders of magnitude smaller than for model A. It is important to realize, that both signs for $`ϵ`$ are generally allowed. In the inflationary case, both the sign and the magnitude of $`ϵ`$ are related to the inflationary slow-roll parameters $`ϵ_{\mathrm{infl}}`$ and $`\eta _{\mathrm{infl}}`$ (Gangui et al., 1994; Gangui, 1994; Wang & Kamionkowski, 1999; Gangui & Martin, 1999). In this model, the PS mass-function at $`z=0`$ is practically recovered on all scales, for all $`|ϵ|\begin{array}{c}<\hfill \\ \hfill \end{array}200`$. #### 3.2.2 Quadratic order in $`ϵ`$ For small deviations from Gaussianity, a first-order expansion in $`ϵ`$ is a valid approximation. We give here the expressions for the relevant quantities also to second order, but in any application we will neglect second or higher-order corrections. When expanding the cumulant generator to second order in $`ϵ`$ we obtain that the variance has a contribution $`ϵ^2`$ as in eq. (37), the skewness remains the same as to first-order in $`ϵ`$, the next to leading term being $`ϵ^3`$, but there is a non-vanishing contribution to the kurtosis $`ϵ^2`$, namely $`\mu _{4,R}`$ $`=`$ $`ϵ^2\mu _{4,R}^{(1)}=48ϵ^2\alpha ^2{\displaystyle \frac{d^3k_1}{(2\pi )^3}\frac{d^3k_2}{(2\pi )^3}\frac{d^3k_3}{(2\pi )^3}P_\varphi (k_1)P_\varphi (k_2)P_\varphi (|𝐤_3+𝐤_1|)}`$ (46) $`\times `$ $`\stackrel{~}{F}_R(k_1)\stackrel{~}{F}_R(k_2)\stackrel{~}{F}_R(k_3)\stackrel{~}{F}_R(|𝐤_1+𝐤_2+𝐤_3|)`$ For both models, the leading-order kurtosis parameter $`S_{4,R}=ϵ^2\mu _{4,R}^{(1)}/(\mu _{2,R}^{(1)})^3`$ becomes scale-independent if the power-spectrum is a power-law; therefore on large scales, where the transfer function is unity, $`S_{4,R}`$ becomes scale-independent. ### 3.3 Higher order non-Gaussian contributions What happens to our results if we allow for higher order terms in the original definition of our primordial non-Gaussian field $`\psi `$? Let us modify, for instance, our definition by adding a cubic term, as follows, $$\psi (𝐱)=\alpha \varphi (𝐱)+ϵ(\varphi ^2(𝐱)\varphi ^2)+ϵ^2\beta \varphi ^3(𝐱),$$ (47) where $`\beta `$ is a new independent parameter, which we assume to be of order unity. Such a cubic term might be easily accounted for in the functional integral approach, by using the so-called ‘integration by parts’ relation (e.g. Ramond 1989). For the purpose of the present paper, however, we can account for such a cubic term through the modifications it induces in the lowest-order cumulants. The sub-leading term of the variance would in fact be modified by the addition of $$ϵ^2\mathrm{\Delta }\mu _{2,R}^{(2)}=6ϵ^2\frac{\beta }{\alpha }\varphi ^2\mu _{2,R}^{(1)},$$ (48) which therefore appears as a ‘renormalization’ of the leading-order variance. The sub-leading skewness would also get an extra contribution of the same order $`ϵ^3`$, which we will not write here. The kurtosis, instead, would be modified already to leading order, gaining the extra piece $$ϵ^2\mathrm{\Delta }\mu _{4,R}^{(1)}=\frac{1}{2}ϵ^2\alpha \beta \mu _{4,R}^{(1)},$$ (49) which still has the nice feature of appearing as a renormalization of the previously calculated leading-order kurtosis. At this point we can derive a general <sup>4</sup><sup>4</sup>4The key assumption that allowed us to obtain such a general treatment of a mildly non-Gaussian field is that it can be expanded as a local functional of an underlying Gaussian field. On the other hand, any form of non-locality such that it can be expressed as a convolution in real space can also be handled in a similar way, by suitably modifying our definition of the function $`F_R(𝐱)`$. and self-consistent approximation to the cumulant generator (and hence for both the differential and cumulative probability) up to order $`ϵ^2`$: $$𝒲(\lambda )=\frac{\lambda ^2}{2}\mu _{2,R}^{(1)}ϵ\frac{i\lambda ^3}{6}\mu _{3,R}^{(1)}+ϵ^2\left[\frac{\lambda ^2}{2}(\mu _{2,R}^{(2)}+6\frac{\beta }{\alpha }\varphi ^2\mu _{2,R}^{(1)})+\frac{\lambda ^4}{24}\left(1+\frac{1}{2}\alpha \beta \right)\mu _{4,R}^{(1)}\right].$$ (50) This is the explicit expression for $`𝒲(\lambda )`$; no further terms could appear from higher-order non-Gaussianity to this order in $`ϵ`$. Substituting this expansion in eq. (26) one gets an approximate expression for the PDF as an integral over $`\lambda `$, which is valid for suitably small values of the non-Gaussian parameter $`ϵ`$. For the rest of our calculations, however, we will assume that departures from Gaussianity are small and therefore we will retain only linear-order terms. ## 4 Press-Schechter approach to non-Gaussian density fields To obtain the abundance of dark matter halos as a function of filtering radius $`R`$ (or mass $`MR^3`$) and redshift of collapse $`z_c`$, one should first obtain the conditional probability that the density contrast equals the threshold for collapse $`\delta _c`$, when filtered on scale $`R`$, provided it is below it on any larger scale $`R^{}`$. In the Gaussian case, and for sharp-k-space filter, this problem has been solved by a number of authors (Peacock & Heavens, 1990; Cole, 1991; Bond et al., 1991) by rephrasing it in terms of the problem of barrier first-crossing by a Markovian random walk. In the non-Gaussian case and/or for other types of filters, different techniques might also be useful. In particular, an alternative formulation, originally proposed by Jedamzik (1995) and successively implemented by Yano, Nagashima & Gouda (1996), Nagashima & Gouda (1997) and Lee & Shandarin (1998), allows to reduce the problem to the solution of the integral equation, $$P(>\delta _c|z_c,M)=\frac{1}{\overline{\rho }_{0m}}_0^{\mathrm{}}𝑑M^{}P(M|M^{})M^{}n(M^{},z_c),$$ (51) where $`\overline{\rho }_{0m}`$ is the present-day mean density of non-relativistic matter, $`P(>\delta _c|z_c,M)`$ is the probability that $`\delta _M`$ lies above the threshold $`\delta _c`$ (i.e. the fraction of volume where this happens) at a given redshift $`z_c`$ and $`n(M,z_c)`$ is the required comoving mass-function for halos of mass between $`M`$ and $`M+dM`$ which formed at $`z_c`$. The function $`P(M|M^{})`$ denotes the conditional probability of finding a region with mass $`M`$ overdense by $`\delta _c`$ or more, given that it is included in an isolated region of mass $`M^{}`$ $`(>M)`$. Let us start by first finding an expression for the L.H.S. of this equation, i.e. for the level-crossing probability $`P(>\delta _c|z_c,M)`$. In looking at this problem, it is convenient to think of the density fluctuation as being time-independent while giving a redshift dependence to the collapse threshold $`\delta _c(z_c)\mathrm{\Delta }_c(z_c)/D(z_c\mathrm{\Omega }_{0m},\mathrm{\Omega }_{0\mathrm{\Lambda }})`$; $`\mathrm{\Delta }_c`$ is the linear extrapolation of the over-density for spherical collapse: it is $`1.686`$ in the Einstein-de Sitter case, while it slightly depends on redshift (see Figure 3) for more general cosmologies (e.g. Kitayama & Suto 1996); $`\mathrm{\Omega }_{0\mathrm{\Lambda }}`$ denotes the closure density of vacuum energy today. Here $`D(z\mathrm{\Omega }_{0m},\mathrm{\Omega }_{0\mathrm{\Lambda }})`$ denotes the general expression for the linear growth factor, which depends on the background cosmology: the redshift dependence of $`D(z_c\mathrm{\Omega }_{0m},\mathrm{\Omega }_{0\mathrm{\Lambda }})^1`$ is shown in Figure 3 for three different cosmological models. Using the spherical isothermal collapse model (Gunn & Gott, 1972), it is possible to relate the mass $`M`$ of a dark halo to the Lagrangian (pre-collapse) comoving length $`R`$ (the sphere of radius $`R`$ will give rise to an object that contains the mass M within the virialization radius) (e.g. White, Efstathiou & Frenk 1993), $$R=\frac{2^{1/2}[V_c/100\mathrm{k}\mathrm{m}\mathrm{s}^1]}{\mathrm{\Omega }_{0m}^{1/2}(1+z_c)^{1/2}f_c^{1/6}}$$ (52) where $`R`$ is in units of Mpc h<sup>-1</sup>. Here $`V_c`$ is the circular velocity required for centrifugal support in the potential of the halo, given by $`M=V_c^3/(10GH(z_c))`$, where $`H`$ denotes the Hubble constant, (e.g. Heavens & Jimenez 1999), and $`f_c`$ is the density contrast at virialization of the newly-collapsed object relative to the background. This is adequately approximated by $`f_c=178/\mathrm{\Omega }_m^{0.6}(z_c)`$ (e.g. Eke, Cole & Frenk 1996). In Figure 3 the dependence of (52) on the cosmology and the redshift of collapse is shown. It is clear that this dependence is very weak and can be ignored for our purposes. For a chosen power-spectrum shape and normalization, the dependence on the cosmological model of the PDF and the level excursion probability is therefore confined in $`\delta _c(z_c)`$. In other words, changing the cosmology is essentially equivalent to changing $`z_c`$ according to the right panel of Figure 3. Following the PS formalism, we can write $$P(>\delta _c|z_c,R)=_{\delta _c(z_c)}^{\mathrm{}}𝑑\delta _RP(\delta _R)=_{\delta _c(z_c)}^{\mathrm{}}𝑑\delta _R_{\mathrm{}}^{\mathrm{}}\frac{d\lambda }{2\pi }e^{i\lambda \delta _R+𝒲(\lambda )},$$ (53) which, exchanging the order of integrations and integrating over $`\delta _R`$, yields the exact and general expression $$P(>\delta _c|z_c,R)=\frac{1}{2\pi i}_{\mathrm{}}^{\mathrm{}}\frac{d\lambda }{\lambda }\mathrm{exp}\left[i\lambda \delta _c(z_c)+𝒲(\lambda )\right]+\frac{1}{2}.$$ (54) This is the exact expression for the probability that at redshift $`z_c`$, the fluctuation on scale $`R`$ exceeds some critical value $`\delta _c`$. From this equation it is possible to obtain an approximate expression for the cumulative probability $`P(>\delta _c|z_c,R)`$ by expanding $`W(\lambda )`$ of eq. (27) to a given order in $`ϵ`$. On the other hand, from knowledge of the cumulants of the cosmological density field smoothed on scale $`R`$, up to some order, it is possible to obtain an approximate expression for the level excursion probability $`P(>\delta |R)`$. The integral (54) can be solved analytically by using the saddle-point technique as in (Fry, 1986) and (Lucchin & Matarrese, 1988). We first perform a Wick rotation, $`\lambda i\lambda `$, in the complex $`\lambda `$ plane, to get rid of the oscillatory behaviour of the integrand. This gives $$P(>\delta _c|z_c,R)=\frac{1}{2\pi i}_i\mathrm{}^i\mathrm{}\frac{d\lambda }{\lambda }\mathrm{exp}\left[\lambda \delta _c(z_c)+(\lambda )\right]+\frac{1}{2},$$ (55) with $`(\lambda )𝒲(i\lambda )`$. Next, let us introduce the effective action $`G(\delta _{\mathrm{eff}})`$, defined as the Legendre transform of $``$, namely $`G(\delta _{\mathrm{eff}})=\lambda \delta _{\mathrm{eff}}(\lambda )`$, with $`\delta _{\mathrm{eff}}=d/d\lambda `$. From the definition of $``$ one has $$P(>\delta _c|z_c,R)=\frac{1}{2\pi i}_{G^{}=i\mathrm{}}^{G^{}=i\mathrm{}}𝑑\delta _{\mathrm{eff}}\frac{G^{\prime \prime }(\delta _{\mathrm{eff}})}{G^{}(\delta _{\mathrm{eff}})}\mathrm{exp}\left[\left(\delta _{\mathrm{eff}}\delta _c(z_c)\right)G^{}(\delta _{\mathrm{eff}})G(\delta _{\mathrm{eff}})\right]+\frac{1}{2}.$$ (56) For large thresholds, $`\delta _c(z_c)1`$, the above integral is dominated by stationary points of the exponential. These occur at $`G^{\prime \prime }(\delta _{\mathrm{eff}})(\delta _c(z_c)\delta _{\mathrm{eff}})=0`$, i.e. at $`\delta _{\mathrm{eff}}=\delta _c(z_c)`$, since $`G^{\prime \prime }(\delta _{\mathrm{eff}})=(d^2/d\lambda ^2)^1>0`$ \[as $``$ must be a convex function of its argument (e.g. Fry 1985)\]. A saddle-point evaluation of this integral then gives (Lucchin & Matarrese, 1988) $$P(>\delta _c|z_c,R)\frac{1}{\sqrt{2\pi }}\frac{(G^{\prime \prime })^{1/2}}{G^{}}\mathrm{exp}(G)|_{\delta _{\mathrm{eff}}=\delta _c(z_c)}.$$ (57) Note that the remaining $`1/2`$ term on the r.h.s. of equation (56) has been exactly canceled by the pole residual of the integral at $`G^{}(\delta _{\mathrm{eff}})=0`$. The latter formula provides an accurate approximation of the level-crossing probability $`P(>\delta _c)`$, provided the cumulant generator $`(\delta _{\mathrm{eff}})`$ is analytic for all finite values of its argument, which is indeed the case for our approximated expression in eq. (50), but certainly not for the exact form of eq. (27). The validity of our last result will then rely on the smallness of the non-Gaussianity parameter $`ϵ`$. It is important to notice here that since - as it is clear from Figure 1 \- the skewness parameter is not small everywhere, we expect the PDF to be very sensitive to small $`ϵ`$ at least for model A. The approximate form of the effective action resulting from Legendre transforming eq. (50) is $$G(\delta _{\mathrm{eff}})\frac{\delta _{\mathrm{eff}}^2}{2\sigma _R^2}\left\{1ϵ^2\left(6\frac{\beta }{\alpha }\varphi ^2+\frac{\mu _{2,R}^{(2)}}{\sigma _R^2}\right)\frac{S_{3,R}}{3}\delta _{\mathrm{eff}}+\left[\frac{1}{4}S_{3,R}^2\frac{1}{12}\left(1+\frac{1}{2}\alpha \beta \right)S_{4,R}\right]\delta _{\mathrm{eff}}^2\right\}.$$ (58) The level crossing probability $`P(>\delta _c)`$ to second order in $`ϵ`$ can be obtained from (57), with (57) and the following substitution: $$\frac{(G^{\prime \prime })^{1/2}}{G^{}}=\frac{\sigma _R}{\delta _{\mathrm{eff}}}\left[1+\frac{ϵ^2}{24\alpha }\left(72\beta \varphi ^2\sigma _R^2\alpha ^2\beta \delta _{\mathrm{eff}}^2\sigma _R^2S_4^{(2)}+12\alpha \mu _2^{(2)}+3S_3^{(1)}\alpha \delta _{\mathrm{eff}}^2\sigma _R^22\delta _{\mathrm{eff}}^2\sigma _R^2\alpha S_4^{(2)}\right)\right]$$ (59) To linear order in $`ϵ`$ the effective action reads $$G(\delta _{\mathrm{eff}})\frac{\delta _{\mathrm{eff}}^2}{2\sigma _R^2}\left\{1\frac{S_{3,R}}{3}\delta _{eff}\right\}.$$ (60) Note that this function is only convex for $`\delta _{\mathrm{eff}}<1/S_{3,R}^{(1)}`$, so that our result can be consistently applied only as long as $`\delta _c(z_c)<1/S_{3,R}^{(1)}`$. We finally obtain $$P(>\delta _c|z_c,R)\frac{1}{\sqrt{2\pi }}\frac{\sigma _R}{\delta _c(z_c)}\mathrm{exp}\left[\frac{1}{2}\frac{\delta _c^2(z_c)}{\sigma _R^2}\left(1\frac{S_{3,R}}{3}\delta _c(z_c)\right)\right].$$ (61) Note that, for the Gaussian case, $`S_{3,R}=0`$, this formula corresponds to the well-known asymptotic behaviour of the complementary error function valid where $`\delta _c2\sigma `$. In particular, for $`S_{3,R}=0`$, and $`z=0`$ this approximation introduces an error of $`20\%`$ at $`R10`$ Mpc $`h^1`$ and an error smaller than $`5\%`$ for $`R>20h^1`$ Mpc. For $`\delta _c(z_c)1/S_{3,R}`$, i.e. for $`R\begin{array}{c}<\hfill \\ \hfill \end{array}10`$ Mpc and/or $`ϵ\begin{array}{c}>\hfill \\ \hfill \end{array}10^3`$ for model A, the integral (54) has to be performed numerically. To first order in $`ϵ`$ eq. (54) becomes: $$P(>\delta _cz_c,R)=\frac{1}{2}\frac{1}{\pi }_0^{\mathrm{}}\frac{d\lambda }{\lambda }\mathrm{exp}(\lambda ^2\sigma _R^2/2)\mathrm{sin}(\lambda \delta _c+\lambda ^3\mu _{3,R}/6).$$ (62) In Figure 4 we show the result of the numerical integration for model A for masses in the range $`10^810^{12}M_{}`$, redshift $`z_c=6,8,10`$ and $`ϵ`$ from top to bottom $`ϵ=10^2`$, $`5\times 10^3`$, $`2\times 10^3`$, $`10^3`$, $`5\times 10^4`$, and the Gaussian case ($`ϵ=0`$, solid line). The main effect of the presence of a small non-Gaussianity such as $`ϵ=5\times 10^4`$, is to amplify $`P(>\delta _cz_c,R)`$ by a factor of order 10. For model B instead, as shown in Figure 2, the skewness parameter for the density field is small and has the opposite sign of $`ϵ`$. For this reason the saddle-point approximation (60) works remarkably well on galaxy scales for $`|ϵ|\begin{array}{c}<\hfill \\ \hfill \end{array}500`$, if $`z_c\begin{array}{c}<\hfill \\ \hfill \end{array}10`$. Figure 5 shows the excursion set probability for model B for masses in the range $`10^810^{12}M_{}`$, redshift $`z_c=6,8,10`$ and $`ϵ`$ from top to bottom $`ϵ=300,200,100,0,100,200`$. The thick solid line is relative to the Gaussian case ($`ϵ=0`$); curves above that have $`ϵ<0`$, below have $`ϵ>0`$. The fact that $`|ϵ|100`$ is needed to create any sizeable departure from the Gaussian $`P(>\delta _c|M)`$, has an important consequence: all conventional inflationary models induce deviations from Gaussianity such that the skewness parameter for the peculiar gravitational potential $`S_{3,\mathrm{\Phi }}`$ is bound to be $`|S_{3,\mathrm{\Phi }}|\begin{array}{c}<\hfill \\ \hfill \end{array}10`$ (Gangui et al., 1994); since $`S_{3,\mathrm{\Phi }}6ϵ`$, we conclude that, in the context of single-field inflation models the level-crossing probability $`P(>\delta _c|M)`$ is indistinguishable from the Gaussian one at least on clusters scale and below. It is important to stress that for this model, using values of $`ϵ`$ of order unity or even larger does not imply the breakdown of the perturbation expansion upon which our results rely, as the coefficients of that expansion, such as the skewness parameter, keep small even for $`|ϵ|\begin{array}{c}<\hfill \\ \hfill \end{array}100`$. The most complex issue is that of finding a suitable expression for the function $`P(M|M^{})`$ in equation (51) in the general non-Gaussian case and for generic filters. In this context, it is useful to define a ‘fudge factor’ $`f`$ through the equation $$P(M|M^{})\frac{1}{f}\mathrm{\Theta }(M^{}M),$$ (63) with $`\mathrm{\Theta }`$ the Heaviside step function. As shown by Nagashima & Gouda (1997), in the Gaussian case, with sharp-k-space filtering, $`P(M|M^{})`$ reduces to the conditional probability $`P(\delta _M\delta _c|\delta _M^{}=\delta _c)`$, which is easily obtained from a bivariate Gaussian, using Bayes theorem; from this one immediately gets $`f=2`$, as expected (e.g. Peacock & Heavens 1990). In the non-Gaussian case (and/or for different types of filter) the value of $`f`$ should be obtained from its very definition. Recently, Koyama, Soda & Taruya (1999) showed that in generic non-Gaussian models one can still write $$f(M,M^{})\left[_{\delta _c}𝑑\omega p(\delta _M=\omega |\delta _M^{}=\delta _c)\right]^1,$$ (64) with $`p(\delta _M=\omega |\delta _M^{}=\delta _c)p(\delta _M=\omega \delta _c)`$, provided $`M^{}M`$. Obtaining a similar relation for all scales $`M^{}>M`$ is a more complex issue, which would require a separate analysis <sup>5</sup><sup>5</sup>5In the Gaussian case, and for sharp-k-space filter, one finds $`p(\delta _M=\omega |\delta _M^{}=\delta _c)=p(\delta _{\stackrel{~}{M}}=\omega \delta _c)`$, for all $`M^{}>M`$, where $`\stackrel{~}{M}`$ is defined by $`\sigma ^2(\stackrel{~}{M})=\sigma ^2(M)\sigma ^2(M^{})`$. Inserting this result in eq. (64) yields $`f=2`$, independently of the mass (Nagashima & Gouda 1997).. Nonetheless, it is extremely reasonable to expect that this approximation works well at least in so far as the deviation from Gaussianity is weak. One then immediately gets the simple result <sup>6</sup><sup>6</sup>6Note that, if the non-Gaussian PDF depends on the filtering mass only through the variance, i.e. $`P(\delta _M)=\sigma _M^1\mathrm{\Psi }(\delta _M/\sigma _M)`$, as assumed by Willick (1999), then $`P(>0|M)=P(>\delta _c|M=0)`$, which yields an alternative expression for $`f`$, used by some authors (e.g. Robinson & Baker 1999). In the most general case, however, the latter result is not valid. $`f1/P(>0|M)`$. In our case, one can compute the cumulative zero-crossing probability $`P(>0,M)`$ numerically, starting from eq. (62). The comoving mass-function of halos formed at redshift $`z_c`$ can thus be immediately obtained by differentiating the integral equation (51). This allows the standard PS formula to be extended to the non-Gaussian case in the simple form (Lucchin & Matarrese, 1988; Colafrancesco et al., 1989; Chiu et al., 1998; Robinson et al., 1999a, b; Koyama et al., 1999): $$n(M,z_c)=f\frac{3H_0^2\mathrm{\Omega }_{0m}}{8\pi GM}\left|\frac{dP(>\delta _c|z_c,R)}{dM}\right|,$$ (65) where we implicitly assumed that the value of $`f`$ has negligible dependence on $`M`$, The factor $`f`$ should account for the physical constraint that all of the clustered matter in the Universe must be included in bound objects of some mass \[i.e. $`_0^{\mathrm{}}𝑑MMn(M,z_c)=3H_0^2\mathrm{\Omega }_{0m}/8\pi G`$\] and simultaneously solve the cloud-in-cloud problem. For the non-Gaussian models considered here, the factor $`f`$ is not much different from the Gaussian value, provided $`ϵ`$ is small, i.e. for small departures from the Gaussian behaviour. In Figure 8 the fudge factor $`f1/P(>0|M)`$ is plotted as a function of the skewness parameter, using a top-hat filter in real space. The useful range is on the left of the vertical dotted line, that is $`M\begin{array}{c}>\hfill \\ \hfill \end{array}2\times 10^{10}M_{}`$ for model A with $`ϵ=10^3`$ and $`M\begin{array}{c}<\hfill \\ \hfill \end{array}4\times 10^{15}M_{}`$ for model B with $`ϵ=100`$. We can therefore conclude that, for mild non-Gaussianity, the correction to the usual fudge factor $`f=2`$ is negligible when compared to the effect of the skewness term in the exponential, justifying therefore the use of $`f=2`$ throughout. ### 4.1 Model A In model A the skewness parameter $`S_{3,R}`$ is scale-independent on large scales, therefore, as a first step, we can neglect the dependence of $`S_{3,R}`$ on mass. In this case, in the above differentiation, we arrive to a very simple generalization of the PS formula, which has the same functional form, provided one makes the replacement $$\delta _c(z_c)\delta _c(z_c)\left[1\frac{S_{3,R}}{3}\delta _c(z_c)\right]^{1/2}\delta _{}(z_c)$$ (66) and therefore: $$n(M,z_c)f\frac{3H_0^2\mathrm{\Omega }_{0m}}{8\pi GM^2}\frac{\delta _{}(z_c)}{\sqrt{2\pi }\sigma _M}\mathrm{exp}\left[\frac{\delta _{}^2(z_c)}{2\sigma _M^2}\right]\left|\frac{d\mathrm{ln}\sigma _M}{d\mathrm{ln}M}\right|$$ (67) This simple result is completely analogous to that found by Lucchin & Matarrese (1988), for hierarchical-type statistics: the effect of the deviation from primordial Gaussianity is that of shifting the PS characteristic mass to higher or lower masses, depending on whether the skewness is positive or negative. Note that the effect of the non-Gaussian tail becomes more and more important at higher redshifts. As noticed by Fry (1986), ‘the departures from Gaussian can be appreciable while fluctuations are still small’. The assumption that $`S_{3,R}`$ is scale-independent is correct on relatively large scales, where $`R>20h^1`$ Mpc, corresponding to masses $`M>10^{14}M_{}`$. Substitution (66) applied to the standard PS formalism agrees with the numerical results on scales $`R>20h^1`$ Mpc, for $`ϵ\begin{array}{c}<\hfill \\ \hfill \end{array}10^2`$. The dependence on the cosmological model is confined to $`\delta _c(z_c)`$, as discussed before, once the present-day power-spectrum shape and normalization have been chosen, and the normalization $`H_0^2\mathrm{\Omega }_{0m}`$. However from Figure 1 it is clear that the scale dependence of $`S_{3,R}^{(1)}`$ is not negligible for galaxy masses, moreover for $`R\begin{array}{c}<\hfill \\ \hfill \end{array}10`$ i.e. $`M\begin{array}{c}<\hfill \\ \hfill \end{array}10^{13}`$ the saddle-point approximation does not hold. Numerical evaluation of (54) yields $`dP(>\delta _c)/dM`$; the corresponding $`n(M,z_c)`$ is shown in Figure 6 for the same range of masses, redshift of collapse $`z_c`$ and non-Gaussianity parameter $`ϵ`$ as in Figure 4, assuming $`f=2`$, according to the discussion above. ### 4.2 Model B It is clear from the scale dependence of the skewness parameter that, for departures from Gaussianity of the kind of model B, clusters scales are better than galaxy scales to probe primordial non-Gaussianity. It is worth to notice here that also for model B in the absence of the transfer function, $`S_{3,R}`$ is scale independent and indeed the skewness parameter becomes scale-independent on very large scales ($`R\begin{array}{c}>\hfill \\ \hfill \end{array}100h^1`$ Mpc), where the transfer function is unity. On clusters scales $`S_{3,R}`$ is not scale-independent, but for $`ϵ\begin{array}{c}<\hfill \\ \hfill \end{array}200`$ and redshift $`z_c\begin{array}{c}<\hfill \\ \hfill \end{array}2`$ the saddle-point approximation is still valid. For the comoving mass-function of halos formed at redshift $`z_c`$ we obtain: $$n(M,z_c)f\frac{3H_0^2\mathrm{\Omega }_{0m}}{8\pi GM^2}\frac{1}{\sqrt{2\pi }\sigma _M}\left|\frac{1}{2}\frac{\delta _c^2(z_c)}{3\sqrt{1S_{3,M}\delta _c(z_c)/3}}\frac{dS_{3,M}}{d\mathrm{ln}M}+\frac{\delta _{}(z_c)}{\sigma _M}\frac{d\mathrm{ln}\sigma _M}{d\mathrm{ln}M}\right|\mathrm{exp}\left[\frac{\delta _{}^2(z_c)}{2\sigma _M^2}\right]$$ (68) Also in this case the dependence on the cosmological model is confined to $`\delta _c(z_c)`$ and in the normalization $`H_0^2\mathrm{\Omega }_{0m}`$. Figure 7 illustrates the effect on the mass-function on clusters scales for the non-Gaussianity of model B. For two different redshifts of collapse ($`z_c=1`$ and $`2`$) the ratio of the mass-function for model B to the mass-function for a Gaussian field \[$`n(M,z_c)/n_{\mathrm{Gau}}(M,z_c)`$\] is plotted as a function of the mass in solar masses. Also in this case we can assume $`f=2`$. The choice for the $`ϵ`$ parameter is, from top to bottom, $`ϵ=100,50,10`$. It is clear that for high masses one is probing the tail of the distribution, that is most sensitive to departures from Gaussianity. The forthcoming X-ray Multi-Mirror (XMM) galaxy cluster survey (Romer et al., 1999), will allow the number density of clusters to be accurately measured even at $`z\begin{array}{c}>\hfill \\ \hfill \end{array}1`$. In particular the XMM cluster survey will provide a complete survey of clusters with masses $`M\begin{array}{c}>\hfill \\ \hfill \end{array}10^{14}M_{}`$ at $`z<1.4`$ over 800 square degrees. Preliminary calculations suggest that it will be therefore possible to detect $`|ϵ|\begin{array}{c}>\hfill \\ \hfill \end{array}50`$. ## 5 A worked example: galaxies at $`z>5`$ The technique presented in the previous section allows one to measure accurately the amount of non-Gaussianity on a given scale. Although the traditional route has been to use the abundance of clusters, we will illustrate our technique by using the observed abundance of high-$`z`$ galaxies (e.g. Cavaliere & Szalay 1986; Kashlinsky & Jimenez 1997; Peacock et al. 1998). There are some advantages in using high-$`z`$ galaxies: they sample directly the galaxy scale and the objects are always inside virialised halos. One major disadvantage is that the mass is not accurately determined. On the other hand, as larger telescopes, such as NGST, get on line, it will be possible to obtain volume limited samples of of high-$`z`$ galaxies and from high-resolution spectra, dynamical masses will be determined more accurately. As seen from eq. (66), it is clear that the higher the redshift the more sensitive the PDF is to non-Gaussianity. Up to date there are 6 galaxies with spectroscopic redshifts observed between $`z>5`$ to $`z7`$ (Dey et al., 1998; Spinrad et al., 1998; Weymann et al., 1998; Hu et al., 1999; Chen et al., 1999) and in Table 1 we have compiled their most relevant characteristics. These high-$`z`$ galaxies have been found in relatively small areas of the sky: for the purposes of this worked example we will use a scanned area of $`3\times 10^4`$ square degrees for each galaxy <sup>7</sup><sup>7</sup>7Four of the six galaxies considered here have been found in the Hubble Deep field, where the total area scanned is about $``$ 1/800 of a square degree. The other two galaxies have been found in Keck observations, where the area scanned for each of them is about $`7.4\times 10^5`$ square degrees. For this worked example we will therefore make the conservative choice that there is on average one galaxy per $`3\times 10^4`$ square degrees.. As pointed out before, the masses of this high-$`z`$ sample are unknown, but their current obscured star formation rate is known. The first step is to correct the observed star formation rate for the presence of dust. In (Jimenez et al., 1999) a case has been made for the correction factor for the star formation due to the presence of dust in a starburst being $`5`$ (independently of the magnitude of the host galaxy). This is confirmed by the independent arguments presented in Peacock et al. (1999). The corrected star formation rate is presented in Table 1. Since the age of the Universe at $`z6`$ is already about 0.7 Gyr even for EdS, assuming the present age to be 13 Gyr, it is very unlikely that galaxies have been observed just when their stellar populations are born. Note that galaxies are likely to be dust–enshrouded for about 0.02 Gyr (Jimenez et al., 1999) and thus be invisible for about 3% of the Hubble time at $`z6`$. The star formation rate of the galaxy in the past is more likely to be higher (Heavens & Jimenez, 1999) than the present observed one. To account for this we will then assume that the galaxy has been observed when it is 0.1 Gyr old (i.e. about 10% of the Hubble time at $`z6`$) and has been forming stars at the current (constant) observed rate. Using this, we can estimate what the mass in stars in the galaxies (see Table 1) and in dark matter is, using a simple isothermal profile for the halo, i.e. $`M=V_c^3/(10GH(z_c))`$, where $`z_c`$ is the redshift of collapse <sup>8</sup><sup>8</sup>8Although we have assumed that the stellar population of the galaxy has to be about 10% of the Hubble time at the observed redshift, we will conservatively consider that the dark halo has just collapsed at the observed redshift. This is what is expected in hierarchical models, where the stellar population is generally made in previous generations inside smaller haloes. and assuming that $`\mathrm{\Omega }_{0b}=0.015h^2`$. Following Peacock et al. (1998) we will carry out our analysis in terms of $`\sigma _v`$. For an isothermal halo $`\sigma _v=V_c/\sqrt{2}`$. Table 1 shows that $`\sigma _v=70200`$ km s<sup>-1</sup>. The number density of high-$`z`$ galaxies is $`N(\mathrm{\Omega }=1)\begin{array}{c}>\hfill \\ \hfill \end{array}3.6\times 10^3(h^1\mathrm{Mpc})^3`$ \[N($`\mathrm{\Omega }_0=0.3,\mathrm{\Lambda }_0=0.7`$)$`\begin{array}{c}>\hfill \\ \hfill \end{array}8.3\times 10^4(h^1\mathrm{Mpc})^3`$\]. To compare these observations with our theoretical results we integrate the mass-function to obtain the number of galaxies observed per unit volume with redshift of formation $`6<z_c<8`$ and masses $`2\times 10^{10}M_{}<M<4\times 10^{11}M_{}`$. It is worth noting that our mass estimates slightly underpredict the masses that are derived using the width of the $`Ly\alpha `$ line (Dey et al., 1998). Gaussian initial conditions give N($`\mathrm{\Omega }_0=0.3,\mathrm{\Lambda }_0=0.7`$)= $`5.2\times 10^5(h^1\mathrm{Mpc})^3`$, which is about a factor sixteen lower than the observed value. For Model A with $`ϵ=10^3`$, we obtain N($`\mathrm{\Omega }_0=0.3,\mathrm{\Lambda }_0=0.7`$)= $`1.3\times 10^3(h^1\mathrm{Mpc})^3`$, which is much closer to the observed value. As we already noted before, a very small non-Gaussianity parameter (i.e. a very small departure from the Gaussian behaviour) has a dramatic effect on the statistics of high redshift objects. This is, of course, very preliminary. For the tail of the PDF, the mass function is very sensitive to the overdensity threshold, the mass determination and the redshift of formation. In this example we have assumed that the redshift of formation is the redshift at which the object has been observed. This is of course quite a conservative assumption, yielding a lower limit on the non-Gaussianity parameter. It has been shown that the PS algorithm with a fixed threshold $`\delta _c`$ does not reproduce very accurately N-body simulation results (e.g. Sheth & Tormen 1999). However, the deviations on the mass scales considered here are never larger than a factor of a few, well below the uncertainty due to the mass determination (see below). PS predictions and N-body results can be reconciled by making the threshold for collapse scale-dependent (e.g. Sheth & Tormen 1999; Sheth, Mo & Tormen 1999). This will change the absolute prediction on the number density of objects but will not strongly affect the ratio between the Gaussian and non-Gaussian number density predictions. The spherical collapse assumption underlying the PS algorithm is the main reason why theory does not reproduce accurately N-body results: the introduction of the ellipsoidal collapse improve sensibly the agreement (Lee & Shandarin 1998). In a subsequent paper we will address the issue of triaxial collapse for non-Gaussian initial conditions. Maybe the most pressing issue is that of the mass determination of high-redshift objects. For example, if the mass estimate for the galaxies in the worked example is wrong by a factor of 4, the number density would change by a factor of 20 (at $`z_c=6`$), and agreement between predictions and observations could be obtained without resorting to $`ϵ0`$. Indeed, better observations will improve the determination of the mass of the dark halos and of the number density of high redshift galaxies. A measurement of $`ϵ`$ and thus a determination of the amount of primordial non-Gaussianity (if any!) present on galaxy scales will then become possible. ## 6 Conclusions We have shown that it is possible to test whether initial conditions were Gaussian, by looking at the abundance of galaxies and clusters at high redshift. Small deviation from Gaussianity have a deep impact on those statistics which probe the tail of the distribution: for this reason the abundance of high redshift objects such as galaxies at $`z\begin{array}{c}>\hfill \\ \hfill \end{array}5`$ and clusters at $`z\begin{array}{c}>\hfill \\ \hfill \end{array}1`$, is very sensitive to small departures from the Gaussian behaviour. We used a parameterization of primordial non-Gaussianity that covers a wide range of physically motivated models: the non-Gaussian field is given by a Gaussian field plus a term proportional to the square of a Gaussian field, the proportionality constant being the non-Gaussianity parameter. We applied this model to the primordial density fluctuation field and to the gravitational potential fluctuation field, and we assumed that the departures from the Gaussian behaviour are small. Given this parameterization of non-Gaussianity, we were able to obtain an analytic expression for the probability density function (PDF) of the smoothed density field. This approach is somewhat different from what is found in the literature (e.g. Robinson et al. 1999a,b), since we do not assume a particular form for the non-Gaussian PDF, but we derive it, given a parameterization of the primordial non-Gaussianity. We then introduced a generalized version of the PS approach valid in the context of non-Gaussian initial conditions, also tackling the cloud-in-cloud problem. This enabled us to relate, analytically, the non-Gaussianity parameter to the number of high-redshift objects. The strength of our method is that we are able to properly take into account the smoothing scale, or mass, dependence of the probability distribution function, and that our results are analytic. In looking at the likelihood of rare events for a non-Gaussian density field, one is probing the tail of the distribution, and analytical results are not only elegant but also extremely important. The main technical results of the present paper can be summarized as follows: a) Eq. (27) is the exact analytic form of the cumulant generator for our non-Gaussian models: it can be used to generate cumulants of any order by simple differentiation. b) Eq. (54): this equation gives an exact and general analytic expression for the cumulative probability that, at a given redshift $`z_c`$, the fluctuations on scale $`R`$ exceed some critical value $`\delta _c`$. From this expression it is possible to obtain the cumulative probability to any order by expanding the cumulant generator (36) with (40) and (41) as we did in (50). c) Eq. (61): it is an analytic expression for the cumulative probability valid for small deviations from the Gaussian behaviour. d) Eq. (68), with the definition (66): it is an analytic expression for the comoving mass-function of halos obtained within the PS approach. For a given power spectrum the dependence on the cosmological model is straightforward: it is confined to the normalization factor $`H_0^2\mathrm{\Omega }_{0m}`$ and the threshold $`\delta _c`$. Still an important issue remains unsolved: how to accurately determine the mass and the formation redshift of observed high-redshift objects in order to put tight constraints on primordial non-Gaussianity. This issue will be addressed in a forthcoming paper. LV acknowledges the support of a TMR grant. LV and RJ thank A. Berera, A. Heavens and J. Peacock for stimulating discussions and the anonymous referee for useful comments. SM thanks K. Jedamzik, H. J. Mo and R. Sheth for useful discussions. Figure 1: Model A variance, skewness ($`\mu _{3,R}^{(1)}`$) and skewness parameter ($`S_{3,R}^{(1)}`$) as a function of the radius of the top-hat window in real space and of the mass. To produce this plot we assumed: $`\alpha =1`$, a scale-invariant initial power-spectrum $`P_\varphi k`$, CDM transfer function, $`\mathrm{\Omega }_{0m}=0.3`$, normalized to produce $`\sigma _8(z=0)=1`$ (more details in the text). Figure 2: Model B variance, skewness ($`\mu _{3,R}^{(1)}`$) and skewness parameter ($`S_{3,R}^{(1)}`$) as a function of the radius of the top-hat window in real space and of the mass. To produce this plot we made the same assumptions as for Figure 1 except that now $`P_\varphi k^3`$. Figure 3: Left panel: The redshift dependence of the threshold for collapse $`\mathrm{\Delta }_c`$ in different cosmologies. It is clear that the dependence on redshift and on cosmology is weak. This justifies to assume $`\mathrm{\Delta }_c`$=1.687 throughout. Middle panel: Dependence of the relation between dark mass and radius on the cosmology and the redshift of collapse $`z_c`$. The solid lines denote an Einstein de Sitter Universe, the dashed lines a flat model with $`\mathrm{\Omega }_{0m}=0.3`$ and $`\mathrm{\Lambda }_0=0.7`$ and the dotted lines an open model with $`\mathrm{\Omega }_{0m}=0.3`$. The dependence is very weak and can be safely ignored. Right panel: The redshift dependence of the inverse of the linear fluctuation growth for different cosmologies. Figure 4: Model A $`P(>\delta _cz_c,R)`$ for $`z_c=6`$ (left panel), $`z_c=8`$ (center), $`z_c=10`$ (right panel) and $`ϵ=10^2`$, $`5\times 10^3`$, $`2\times 10^3`$, $`10^3`$, $`5\times 10^4`$. The thick solid line is relative to the Gaussian case. It is clear that the main effect of the presence of a small nonzero $`ϵ`$ is to boost $`P(>\delta _cz_c,R)`$ by at least a factor $`10`$. Figure 5: Model B $`P(>\delta _cz_c,R)`$ for $`z_c=6`$ (left panel), $`z_c=8`$ (center), $`z_c=10`$ (right panel) and $`ϵ=300`$, $`200`$, $`100`$, $`0`$, $`100`$, $`300`$ (from top to bottom). The thick solid line is relative to the Gaussian case. It is clear that $`ϵ_B`$ of model B need to be $`ϵ_A`$ of model A to show a noticeable difference from the Gaussian case. Moreover $`ϵ_B<0`$ is needed in order to enhance the non-Gaussian tail of $`P(>\delta _cz_c,R)`$. Figure 6: Model A: The comoving mass-function of halos formed at redshift 6 (left panel) 8 (center), 10 (right panel), for different values of the non-Gaussian parameter $`ϵ`$, as in Figure 4; from top to bottom: $`ϵ=10^2`$, $`5\times 10^3`$, $`2\times 10^3`$, $`10^3`$, $`5\times 10^4`$. The thick solid line is relative to the Gaussian case. Figure 7: Model B: The non-Gaussian effect on the mass-function on clusters scales. For two different redshift of collapse ($`z_c=1`$ left and $`z_c=2`$ right) the ratio of the mass-function for model B to the mass-function for a Gaussian field \[$`n(M,z_c)/n_{\mathrm{Gau}}(M,z_c)`$\] is plotted as a function of mass. The choice for the $`ϵ`$ parameter is, from top to bottom, $`ϵ=100,50,10`$. It is clear that for high masses one is probing the tail of the distribution, that is most sensitive to departures from Gaussianity. Figure 8: The fudge factor $`f1/P(>0|M)`$ as a function of the skewness parameter. The value $`f=2`$ is exact for the Gaussian case with sharp-k-space filter; the solid line shows the value of $`f`$ as a function of the skewness parameter $`S_{3,R}`$. The useful range is on the left of the vertical dotted line, i.e. $`M\begin{array}{c}>\hfill \\ \hfill \end{array}2\times 10^{10}M_{}`$ for model A with $`ϵ=10^3`$ and $`M\begin{array}{c}<\hfill \\ \hfill \end{array}4\times 10^{15}M_{}`$ for model B with $`ϵ=100`$. We can conclude that, for mild non-Gaussianity, the correction to the usual fudge factor is negligible, justifying therefore the use of $`f=2`$ throughout.
warning/0001/quant-ph0001099.html
ar5iv
text
# PHYSICAL MODEL OF SCHRODINGER ELECTRON. HEISENBERG CONVENIENT WAY FOR DESCRIPTION OF ITS QUANTUM BEHAVIOUR ## 1 Introduction We assume that the vacuum fluctuations (VcmFlcs) through zero-point quantum electromagnetic field (QntElcMgnFld) perform an important role in a behaviour of the micro particles (MicrPrts). As thing turned out that if the Brownian stochastic motion (BrnStchMtn) of some classical micro particle (ClsMicrPrt) is a result of fluctuating deviations of averaged values of all having an effect forces on a ClsMcrPrt, coming from many molecule blows from a surround environment, then the quantized stochastic dualistic wave-particle behaviour of every QntMicrPrt is a result of the continuous uncontrolled electromagnetic interaction (ElcMgnIntAct) between its well spread (WllSpr) elementary electric charge (ElmElcChrg) of the charged one (a Schrodinger’s electron (SchEl)) or its magnetic dipole moment (MgnDplMmn) for uncharged one (as a neutron), and the averaged electric intensity for charged MicrPrts or the averaged magnetic intensity for uncharged one, of the resultant quantized electromagnetic field (QntElcMgnFld) of all stochastic virtual photons (StchVrtPhtns), excited within the FlcVcm and existing within its neighborhood, which exercises a very power influence on its state and behaviour. Consequently the continuously scattering of the well spread (WllSpr) elementary electric charge (ElmElcChrg) of the SchEl on the StchVrtPhtns at its creation powerfully broken the smooth thin line of the classical trajectory of many short and very disorderly orientated small lines and the powerfully its interaction (IntAct) with the electric intensity (ElcInt) or magnetic intensity (MgnInt) of the resultant QntElcMgnFld of the existing StchVrtPhtns forced it to make the circular harmonic oscillations with various radii and the centers, lying over the small disordered lines. In a result of this complicated motion the narrow smooth line of the classical trajectory is turned out into some wide rough cylindrically spread path of the QntMcrPrt. Although in further we will give the necessary calculations, we wish to repeated, that as a result of the continuously scattering of the QntMicrPrt on the StchVrtPhtns at their creations the smooth thin line of the classical trajectory is turned out into powerfully often broken of many small and very disorderly orientated short lines. The uninterrupted ElcMgnIntAct of the ElmElcChrg or of the MgnDplMmn of the QntMicrPrt with the ElcInt or the MgnInt of the resultant QntElcMgnFld of the StchVrtPhtns, existent within the fluctuating vacuum (FlcVcm) between two consecutive scatterings forced the QntMicrPrt to carry out the stochastic circular oscillation motion, which exercise an influence of its behavior within a neighborhood of the smooth classically line into the cylindrically spread by different radii wide path. It isn’t allowed us to forget that the broken of the smooth thin line of very short and very disorderly orientated small line is a result of its continuously scattering on the StchVrtPhtns, over which are found the centers of the forced stochastic circular oscillation motion of the QntMicrPrt, owing a result of the ElcMgnIntAct of its WllSpr ElmElcChrg and MgnDplMmn with the intensities of the resultant electric field (RslElcFld) and resultant magnetic field (RslMgnFld) of the stochastic virtual photons (StchVrtPhtns). The WllSpd ElmElcChrg of the SchEl is moving at its circular oscillations of different radii within the flats, which are perpendicular to the very short and very disorderly orientated small lines, obtained in a result of its continuously scattering on the StchVrtPhtns, at its Furtian quantized stochastic circular harmonic oscillation motion through the fluctuating vacuum (FlcVcm). Therefore in our transparent survey about the physical model (PhsMdl) of the nonrelativistic quantized SchrEl one will be regarded as some WllSpr ElmElcChrg), participating simultaneously in two different motions: A) The classical motion of a classical Lorentz’ electron (LrEl) along an well contoured smooth thin trajectory, realized in a consequence of some known interaction (IntAct) of its over spread (OvrSpr) ElmElcChrg, MgnDplMnt or bare mass with the intensity of some external classical fields (ClsFlds) as in the Newton nonrelativistic classical mechanics (NrlClsMch) and Maxwell nonrelativistic classical electrodynamics (ClsElcDnm). B) The isotropic three-dimensional nonrelativistic quantized (IstThrDmnNrlQnt) Furthian stochastic boson circular harmonic oscillations motion (FrthStchBsnCrcHrmOscMtn) of the SchEl as a natural result of the permanent ElcIntAct of its WllSpr ElmElcChrg with the ElcInt of the resultant QntElcMgnFld of a large number StchVrtPhtns. This ElcIntAct between the WllSpr ElmElcChrg and the FlcVcm (zero-point ElcMgnFld) is generated by dint of StchVrtPhtns exchanged between the fluctuating vacuum (FlcVcm) and the WllSpr ElmElcChrg during a time interval of their life. As soon as this Furthian quantized stochastic wave-particle behaviour of the SchEl is very similar to known Brownian classical stochastic behaviour of the ClsMacrPrt, therefore the QntMicrPrt cannot has the classical sharp contoured smooth and thin trajectory but has a cylindrical broad rough path, obtained as a sum of circular oscillations motions of different radii and centers, lying on accidental broken short lines, strongly disordered within a space. Hence the often broken trajectory of the moving QntMicrPrt present itself a sum of small parts from some circumferences with different radii and centers, lying within flats, which are perpendicular to accidental broken short lines, strongly disordered in space. Therefore in a principle the exact description of the resultant behaviour of the SchEl owing of its joint participation in both mentioned above motions could be done only by means of the NrlQntMch’s and nonrelativistic ClsElcDnm’s laws. It is known of many scientists the existence of three different ways , and , and , for the description of the quantum behaviour of the nonrelativistic SchEl. It is turned out that there is some possibility enough to show by means of the existence intrinsic analogy between the quadratic differential wave equation in partial derives (QdrDfrWvEqtPrtDrv) of Schrodinger and the quadratic differential particle equation in partial derivative (QdrDfrPrtEqtPrtDrv) of Hamilton-Jacoby that the addition of the kinetic energy of the Furthian stochastic boson circular harmonic oscillation of some QntMicrPrt to the kinetic energy of such ClsMacrPrt determines their dualistic wave-particle quantized behaviour. It turns out the stochastic motion over the powerfully break up the sharp contoured smooth thin classical line of the in many shortly and very disorderly (stochastically orientated) small lines. As in such a natural way we have ability enough to obtain the minimal value of the dispersion product, determined with the Heisenberg uncertainty relation. Science there exists an essential analogy between the registration forms of the quadratic differential diffusive equation (QdrDfrDfsEqt) of Focker-Plank for the distribution function $`P(r,t)`$ of a probability density (DstFncPrbDns) of the free Brownian ClsMacrPrt (BrnClsMacrPrt) in a motionless coordinate system in a respect to it and quadratic differential wave equation in partial derivative (QdrDfrWvEqtPrtDrv) of Schrodinger for the orbital wave function (OrbWvFnc) of some free Furthian QntMicrPrt (FrthQntMicrPrt) in a motionless coordinate system in a respect to it we come to an essential conclusion that there are also some possibility enough to describe the quantized stochastic behaviour of the SchEl by means of the analogy between the classical Wiener continual integral and the quantized Feynman continual integral. Feynman has used for the description of transition between two OrbWvFncs of some free FrthQntMicrPrt with different coordinates and times some formula, analogous of such the formula, which early had been used by Einstein , , Smoluchowski and Wiener for the description of same transition between two DstFncsPrbDns of the free BrnClsMacrPrts. In this way we understand why the behaviour of the QntMicrPrt must be described by the OrbWvFnc $`\mathrm{\Psi }`$ , although the behaviour of the ClsMacrPrt may be described by a line. ## 2 Mathematical description of the physical cause ensuring the display of the QntMicrPrt behaviour. The object of this paper is to discuss the fundamental problems of the physical interpretation of the nonrelativistic quantized behaviour of the SchEl and bring to light for understanding the cause,securing the existence of this uncommon state of each the QntMicrPrt. It is necessary to understand why the QntMicrPrt has no classical smooth thin trajectory and why its behaviour must be described by the Heisenberg matrix of the convenient operator way, using the laws of the NrlQntMch and its effective mathematical results. The PhsMdl of the SchEl is built by means of the equation of the forced motion of the dumping classical oscillator under the force action of electric interaction (ElcIntAct) between its WllSpr ElmElcChrg and the ElcInt of the RslQntElcMgnFld of the StchVrtPhtns, created in the FlcVcm. The unusual behaviour of the SchEl may be described by the following motion equation in Maxwell nonrelativistic classical electrodynamics (ClsElcDnm): $$\ddot{r_j}+\omega _o^2r_j=(\frac{e}{m})\{E_j+E_j^i\}=\frac{e}{mC}\frac{A_j}{t}+\frac{2e^2}{3mC^3}\stackrel{\mathrm{}}{r_j},$$ (1) where $`E_j^i`$ and $`E_j`$ denote the ElcInt of both the ElcFld $`E_j^i`$ of radiative friction, that is to say of LwEng unemitted longitudinal (Lng) VrtPhtn (VrtLngPht), and ElcInt $`E_j`$ of the LwEng-VrtPhtn in the FlcVcm. In accordance of the relation (1) the ElcInt $`E_j`$ of an external QntElcMgnFld may be described by means of its $`A_j`$, having the following analytical presentation: $$A_j=\frac{i}{L}\underset{q}{}\sqrt{\frac{2\pi \mathrm{}C}{Lq}}I_{jq}\left[a_{jq}^+e^{i(t\omega qr)}a_{jq}e^{i(t\omega qr)}\right],$$ (2) Indeed the ElcInt $`E_j`$ of StchVrtFtn could be obtained by taking of a particle derivative of the expression (2) relatively for the $$E_j=\frac{1}{L}\underset{q}{}\sqrt{\frac{2\pi \mathrm{}\omega }{L}}I_{jq}\left[a_{jq}^+e^{i(t\omega qr)}+a_{jq}e^{i(t\omega qr)}\right],$$ (3) There is a necessity to note here that we have exchanged the signs in eqs. (2) and (3). Indeed, in order to get the necessary correspondences between operator expressions of the $`\widehat{p}_j`$ and $`\widehat{A}_j`$, it is appropriate to use the sign (-) in the eq.(2) and the sign (+) in an eq.(3). The helpful of this exchange of signs will be letter seen in following expressions (5) of $`\widehat{r_j}`$ and (6) of $`\widehat{p_j}`$. Hence by substituting the eq.(2) in the eq.(1) and transposition of same term in its left-hand side one can obtain motion equation in Lorentz-Abrahams nonrelativistic presentation (LAP): $$\ddot{r_j}\tau \stackrel{\mathrm{}}{r_j}+\omega _o^2r_j=(\frac{e}{m})E_j,$$ (4) The temporary dependence of $`r_j`$ contains two frequencies $`\omega _o`$ and $`\omega `$. In a spite of $`\omega _o\omega `$, then the very greatest magnitude of the term $`\tau \stackrel{\mathrm{}}{r_j}`$ is -$`\tau \omega _o^2\dot{r_j}`$. Although of that the term $`\tau \stackrel{\mathrm{}}{r_j}`$ still presents itself by -$`\tau \omega ^2\dot{r_j}`$. Indeed, the general solution of eq.(4) is given by sum of the general solution of the homogeneous equation and a particular solution to the inhomogeneous equation. At $`\omega \tau =\frac{2e^2}{3mC^2}\frac{\omega }{C}=\frac{\pi }{3}(\frac{2e^2}{C\mathrm{}})\frac{\mathrm{}}{mC}\frac{2}{\lambda }\mathrm{\hspace{0.17em}1},`$ the general solution of the homogeneous equation has a form of a relaxing oscillation of a frequency $`\omega `$. The particular solution has a form of a forced oscillation of a frequency $`\omega `$. Therefore we may rewrite eq. (4) in the following form : $$\ddot{r_j}+\tau \omega ^2\dot{r_j}+\omega _o^2r_j=(\frac{e}{m})E_j(r,t),$$ (5) From eq.(5) it is easily seen that the motion dumping of the SchEl is caused by well-known Lorentz’ dumping force owing to radiation friction of its moving WllSpr ElmElcChrg. In a rough approximation of the Maxwell nonrelativistic ClsElcDnm the minimum time interval for an emission or absorption of a real photon (RlPhtn) by the WllSpr ElmElcChrg of the SchEl may be evidently determined by the parameter of Lorentz-Abrahams : $$\tau =\frac{2e^2}{3mC^3},$$ (6) The particular solution of the motion eq.(5), describing the forced quantized stochastic circular harmonic motion of the QntMicrPrt, have been written by Welton , Kalitchin and Sokolov and Tumanov , citeAAS by the way of the operator division in the following analytical form : $$\widehat{r_j}=\underset{q}{}\frac{eq}{mL}\sqrt{\frac{2\pi \mathrm{}\omega }{Lq}}I_{jq}\left[\frac{a_{jq}^+\mathrm{exp}\{it\omega iqr\}}{\omega _o^2\omega ^2+i\tau \omega ^3}+\frac{a_{jq}\mathrm{exp}\{it\omega +iqr\}}{\omega _o^2\omega ^2i\tau \omega ^3}\right],$$ (7) $$\widehat{P_j}=i\underset{q}{}\frac{e\omega _o^2}{CL}\sqrt{\frac{2\pi \mathrm{}\omega }{Lq}}I_{jq}\left[\frac{a_{jq}^+\mathrm{exp}\{it\omega iqr\}}{\omega _o^2\omega ^2+i\tau \omega ^3}\frac{a_{jq}\mathrm{exp}\{it\omega +iqr\}}{\omega _o^2\omega ^2i\tau \omega ^3}\right],$$ (8) The analytical presentation (8) of the SchEl’s momentum components have been calculated through using the relation known from Maxwell ClsElcDnm : $$\widehat{P}_j=m\dot{r}_j(\frac{e}{C})\left[A_j+A_j^i\right],$$ (9) Further they have calculated the well-known Heisenberg’s commutation relations (HsnCmtRlts) between the operators of the dynamic variables $`\widehat{r}_j`$ (7) and $`\widehat{P}_j`$ (8) by virtue of the following definition : $$\widehat{P}_j\widehat{r}_k\widehat{r}_k\widehat{P}_ji\mathrm{}\delta _{jk}$$ (10) Since then it is easily to understand by means of the upper account that if the ClsMacrPrt’s motion is occurred along a clear definite smooth thin trajectory in the NrlClsMch, then the QntMicrPrt’s motion is performed in a form of the RndTrmMtn along a pete very small line, stochastically orientated in the space near the clear-cut smooth thin trajectory in the NrlQntMch. As a result of that we can suppose that the QntStchBhv of the QntMicrPrt can be described by means of the following physical quantities in the NrlQntMch : $$r_j=\overline{r}_j+\delta r_j;p_j=\overline{p}_j+\delta p_j;$$ (11) ## 3 Mathematical description of the minimal dispersions of some dynamical parameters of a QntMicrPrt Indeed,because of the eqs.(11) the values of the averaged physical parameters in the NrlQntMch $`p_j^2`$ is different from the values of the same physical parameters in the NrlClsMch $`\overline{p}_j^2`$ as it is seen : $$r_j^2=\overline{r}_j^2+\delta r_j^2;p_j^2=\overline{p}_j^2+\delta p_j^2;$$ (12) In spite of that the averaged value of the orbital (angular) mechanical momentum of the QntMicrPrt has the following value : $$L^2=\underset{j}{}(\overline{L}_j)^2+\underset{j}{}(\delta L_j)^2=(\overline{L}_x)^2+(\delta L_x)^2+(\overline{L}_y)^2+(\delta L_y)^2+(\overline{L}_z)^2+(\delta L_z)^2;$$ (13) or at the $`(\overline{L}_x)^2=\mathrm{\hspace{0.17em}0}`$ and $`(\overline{L}_y)^2=\mathrm{\hspace{0.17em}0}`$ we must obtain : $$L^2=(\overline{L}_z)^2+(\delta L_z)^2+(\delta L_y)^2+(\delta L_x)^2$$ (14) As both the value of the $`(\delta L_x)^2`$ and $`(\delta L_y)^2`$ are equal of the $`\frac{\overline{L}_z\mathrm{}^2}{2}`$ and the value of the $`(\delta L_z)^2`$ is equal of the $`\frac{\mathrm{}^2}{4}`$. Therefore : $$L^2=l^2\mathrm{}^2+l\mathrm{}^2+\frac{\mathrm{}^2}{4}=(l+\frac{1}{2})^2\mathrm{}^2;$$ (15) The realized above investigation assists us to come to the conclusion that the dispersions of the dynamical parameters of the QntMicrPrt are natural results of their forced stochastic oscillation motions along the very small line stochastically orientated in space near to the classical clear-cut smooth thin line of the corresponding dynamical parameters values of the ClsMacrPrt, owing to ElcMgnIntAct of its OvrSpr ElmElcChrg or MgnDplMm with the intensities of the RslElcFld or RslMgnFld of the QntElcMgnFlds of the StchVrtPhtns at its motion through the FlcVcm. It is turned out that the kinetic energy of the IstThrDmnNrlQnt FrthStchBsnCrcHrmOscs, which the QntMicrPrt takes from the FlcVcm, called as its localized energy, one ensures the stability of the SchEl in its ground state in the H-like atom. We have the ability to obtain the minimal value of the dispersion product, determined by the Heisenberg uncertainty relation. In a consequence of what was asserted above in order to obtain the QntQdrDfr WvEqn of Sch we must add to the kinetic energy $`\frac{(_lS_1)^2}{2m}`$ of the NtnClsPrt in the following ClsQdrDifPrtEqt of Hml-Jcb : $$\frac{S_1}{t}=\frac{(_jS_1)^2}{2m}+U;$$ (16) the kinetic energy $`\frac{(_lS_2)^2}{2m}`$ of the BrnClsPrt. In such the natural way we obtain the following analytic presentation of the QntQdrDfrWvEqt of Sch : $$\frac{S_1}{t}=\frac{(_jS_1)^2}{2m}+\frac{(_jS_2)^2}{2m}+U;$$ (17) The purpose of our investigation in henceforth is to obtain the eq. (17) by means of physically obvious and mathematically correct proof. Therefore we could desire a voice of a supposition that all uncommon ways of the SchEl’s behaviour in the NrlQntMch or of other QntMicrPrts in the micro world are natural consequences of unconstrained stochastic joggles on account of continuously accidently exchanges of LwEnr-StchVrtPhtn between its WllSpr ElmElcChrg and the VcmFlc. In consequence of the absence of SchEl’s trajectory within the NrlQntMch as within the NrlClsMch and the stochastical character of its random trembling motion together with the probably interpretation of the SchEl’s OrbWvFnc module square are naturally consequences of the continuous ElcMgnIntAct between the SchEl’s WllSpr ElmElcChrg and EfcElcInt $`E_j`$ of existent LwEnr-VrtPhtns, stochastically generated by fluctuating energy within FlcVcm through continuous incident exchange of LwEnr-StchVrtPhtns, which are either emitted or adsorbable by either the VcmFlcs or the Schel’s Wllspr ElmElcChrg. Really, a deep understanding of the physics of the random trembling motion, in accordance with the description of the Brownian stochastic behaviour of BrnCslPrts we can determine both as the value $`V^{}`$ of the SchEl’s velocity before the moment $`t`$ of the scattering time of some LwEnr-StchVrtPhtns from its WllSpr ElmElcChrg, so the value $`V^+`$ after the same moment $`t`$ of the scattering time by means of the following definitions : $$V_j^{}=lim_{\mathrm{\Delta }to}\left\{\frac{r(t)_jr(t\mathrm{\Delta }t)_j}{\mathrm{\Delta }t}\right\}=(V_jiU_j);$$ (18) $$V_j^+=lim_{\mathrm{\Delta }to}\left\{\frac{r(t+\mathrm{\Delta }t)_jr(t)_j}{\mathrm{\Delta }t}\right\}=(V_j+iU_j);$$ (19) In addition we may determine two new velocities $`V_j`$ and $`U_j`$ by dint of the following equations : $$2V_j=V_j^++V_j^{}\mathrm{and}2iU_j=V_j^+V_j^{},$$ (20) In conformity with the eq.(20) it is obviously followed that the current velocity $`V`$ describes the regular drift of the SchEl and the osmotic velocity $`U`$ describes its nonrelativistic quantized stochastic bozon oscillations. Afterwards by virtue of the well-known definition equations : $$2mV_j=m(V_j^++V_j^{})=\mathrm{\hspace{0.17em}2}_jS_1$$ (21) and $$2imU_j=m(V_j^+V_j^{})=\mathrm{\hspace{0.17em}2}i_jS_2$$ (22) one can obtain the following presentation of the SchEl’s OrbWvFnc $`\psi (r,t)`$ : $$\psi (r,t)=\mathrm{exp}\{i\frac{S_1}{\mathrm{}}\frac{S_2}{\mathrm{}}\}=B\mathrm{exp}\{i(\frac{S_1}{\mathrm{}}\}$$ (23) It is easily to verify the results (20), (21) (22). In an effect ones may be obtained by means of the following natural equations : $$mV_j^+\psi (r,t)=i\mathrm{}_j\mathrm{exp}\{\frac{iS_1}{\mathrm{}}\frac{S_2}{\mathrm{}}\}=(_jS_1+i_jS_2)\psi (r,t)$$ (24) and $$mV_j^{}\psi (r,t)^+=+i\mathrm{}_j\mathrm{exp}\{\frac{iS_1}{\mathrm{}}\frac{S_2}{\mathrm{}}\}=(_jS_1i_jS_2)\psi (r,t)^+$$ (25) Indeed, $`2mV_j=m(V_j^++V_j^{})=`$ $`\left\{(_jS_1+i_jS_2)+(_jS_1i_jS_2)\right\}\mathrm{or}2mV_j=\mathrm{\hspace{0.17em}2}_jS_1`$ (26) and $`2imU_j=m(V_j^+V_j^{})=`$ $`\left\{(_jS_1+i_jS_2)(_jS_1i_jS_2)\right\}\mathrm{or}2imU_j=\mathrm{\hspace{0.17em}2}i_jS_2`$ (27) In consequence we could assume that the module square of the SchEl’s OrbWvFnc $`\psi (r,t)`$ describes the probability density of its location close by the space point $`r`$ at the time moment $`t`$ in the good light of our obvious interpretation. Further in order to obtain the partial differential equation of the continuity we are going to calculate one by virtue of its well-known definitions : $`{\displaystyle \frac{\psi ^2}{t}}+_j(V_j^+\psi ^2)={\displaystyle \frac{(\mathrm{exp}\{\mathrm{\hspace{0.17em}2}\frac{S_2}{\mathrm{}}\})}{t}}+_j\left[(_j{\displaystyle \frac{S_1}{m}}+i_j{\displaystyle \frac{S_2}{m}})\mathrm{exp}\{\mathrm{\hspace{0.17em}2}{\displaystyle \frac{S_2}{\mathrm{}}}\}\right]=`$ $`\left[{\displaystyle \frac{2}{\mathrm{}}}{\displaystyle \frac{S_2}{t}}+{\displaystyle \frac{1}{m}}(_j)^2S_1+{\displaystyle \frac{i}{m}}(_j)^2S_2{\displaystyle \frac{2}{m\mathrm{}}}_jS_1_jS_2{\displaystyle \frac{2i}{\mathrm{}}}_jS_2_jS_2\right]\left[\mathrm{exp}\{\mathrm{\hspace{0.17em}2}{\displaystyle \frac{S_2}{\mathrm{}}}\}\right]`$ (28) $`{\displaystyle \frac{\psi ^2}{t}}+_j(V_j^{}\psi ^2)={\displaystyle \frac{(\mathrm{exp}\{\mathrm{\hspace{0.17em}2}\frac{S_2}{\mathrm{}}\})}{t}}+_j\left[(_j{\displaystyle \frac{S_1}{m}}i_j{\displaystyle \frac{S_2}{m}})\mathrm{exp}\{\mathrm{\hspace{0.17em}2}{\displaystyle \frac{S_2}{\mathrm{}}}\}\right]=`$ $`\left[{\displaystyle \frac{2}{\mathrm{}}}{\displaystyle \frac{S_2}{t}}{\displaystyle \frac{1}{m}}(_j)^2S_1{\displaystyle \frac{i}{m}}(_j)^2S_2{\displaystyle \frac{2}{m\mathrm{}}}_jS_1_jS_2+{\displaystyle \frac{2i}{\mathrm{}}}_jS_2_jS_2\right]\left[\mathrm{exp}\{\mathrm{\hspace{0.17em}2}{\displaystyle \frac{S_2}{\mathrm{}}}\}\right]`$ (29) With the purpose to calculate the last expressions of the continuity equations (3) and (3) we are going to turn the expression (23) of the SchEl’s OrbWvFnc $`\psi (r,t)`$ in the quadratic differential wave equation in partial derivatives of Schrodinger : $$i\mathrm{}\frac{\psi (r,t)}{t}=\frac{\mathrm{}^2}{2}\frac{(_j)^2}{m}\psi (r,t)+U(r,t)\psi (r,t)$$ (30) Further we are able to obtain the following result : $`\left({\displaystyle \frac{S_1}{t}}i{\displaystyle \frac{S_2}{t}}\right)\psi (r,t)=`$ $`\left\{{\displaystyle \frac{(_jS_1)^2}{2m}}{\displaystyle \frac{(_jS_2)^2}{2m}}+{\displaystyle \frac{\mathrm{}}{2m}}(_j)^2S_2i{\displaystyle \frac{\mathrm{}}{2m}}(_j)^2S_1+{\displaystyle \frac{i}{m}}_jS_1_jS_2+U(r,t)\right\}\psi (r,t)`$ (31) As there exist both the real and imaginary parts in the complex valued eq. (3) , it is obviously that from this one follows two quadratic differential equations in partial derivatives : $$\frac{S_2}{t}=\frac{\mathrm{}}{2m}(_j)^2S_1\frac{1}{m}(_jS_1)(_jS_2)$$ (32) and $$\frac{S_1}{t}=\frac{1}{2m}(_jS_1)^2\frac{1}{2m}(_jS_2)^2+\frac{\mathrm{}}{2m}(_j)^2S_2+U(r,t)$$ (33) Inasmuch as it is well-known from the NrlQntMch the continuity partial differential equation can be obtained by means of the eqs.(32), (21 and (23) in the following form : $`{\displaystyle \frac{\left|\psi \right|^2}{t}}+_j\left(V_j\left|\psi \right|^2\right)={\displaystyle \frac{\mathrm{exp}\{\mathrm{\hspace{0.17em}2}\frac{S_2}{\mathrm{}}\}}{t}}+{\displaystyle \frac{1}{m}}_j\left(_jS_1\mathrm{exp}\{\mathrm{\hspace{0.17em}2}{\displaystyle \frac{S_2}{\mathrm{}}}\}\right)=\mathrm{\hspace{0.17em}0};`$ (34) Thence the eq.(3) and eq.(3) can be simplified by means of the eqs.(34) and (32). In a result of such substitutions the following continuity partial differential equations could be obtained : $$\frac{\left|\psi \right|^2}{t}+_j\left(V_j^+\left|\psi \right|^2\right)=\frac{i}{m}_j\left(_jS_2\mathrm{exp}\{2\frac{S_2}{\mathrm{}}\}\right)$$ (35) $$\frac{\left|\psi \right|^2}{t}+_j\left(V_j^{}\left|\psi \right|^2\right)=\frac{i}{m}_j\left(_jS_2\mathrm{exp}\{2\frac{S_2}{\mathrm{}}\}\right)$$ (36) In order to calculate the value of the expressions in the brackets in the right-hand side of the eqs.(35) and (36) we will determine the relation between the values of both integrals : $$_{V_R}_{j}^{}{}_{}{}^{2}S_2\mathrm{exp}\{2\frac{S_2}{\mathrm{}}\}dV\mathrm{and}_{V_R}(_jS_2)^2\mathrm{exp}\{2\frac{S_2}{\mathrm{}}\}𝑑V$$ (37) The first integral in (37) may be calculated through integration by parts. In this easily way we could obtain : $`{\displaystyle _{V_R}(_j)^2S_2\mathrm{exp}\{2\frac{S_2}{\mathrm{}}\}𝑑V}={\displaystyle _{S_R}_jS_2\mathrm{exp}\{2\frac{S_2}{\mathrm{}}\}dS_j}`$ $`{\displaystyle _{S_o}_jS_2\mathrm{exp}\{2\frac{S_2}{\mathrm{}}\}dS_j}+{\displaystyle \frac{2}{\mathrm{}}}{\displaystyle _{V_R}(_jS_2)^2\mathrm{exp}\{2\frac{S_2}{\mathrm{}}\}𝑑V}`$ (38) From above it is evidently that the second two-dimensional integral over the surface $`S_o`$ cannot exist in the case when the integrational domain $`V_R`$ of the three-dimensional integral has the form of one-piece-integrity domain. Indeed, the three-multiple integral in the left-hand side of eq. (3) has an integration domain of the volume $`V_R`$,then the both two-multiple integrals (the first and second ones on the right handside of the same equation) have a integration domain in form of surface of same volume (the outer skin $`S_R`$ and the inter skin $`S_o`$ of the volume $`V_R`$). Inasmuch as we don’t take into account the creation and annihilation of the FrthQntMicrPrt in the NrlQntMch, than the SchEl’s OrbWvFnc $`\psi (r,t)`$ may have no singularity within the volume $`V_R`$. Therefore the three-multiple integrals have the one-piece integrity domain of an integration without its inter skin surface $`S_o`$. Hence it is easily seen that both two-multiple integrals are canceled in the case when R go to $`\mathrm{}`$ and at the absence of any kind of singularity in the SchEl’s OrbWvFnc. Consequently eq.(37) becomes the form : $`{\displaystyle _V_{\mathrm{}}(_j)^2S_2\mathrm{exp}\{2\frac{S_2}{\mathrm{}}\}𝑑V}={\displaystyle \frac{2}{\mathrm{}}}{\displaystyle _U_{\mathrm{}}(_jS_2)^2\mathrm{exp}\{2\frac{S_2}{\mathrm{}}\}𝑑V}`$ (39) Then in a result of the existence of the eqt.(39) we may suppose the existence of the following equations between the values of both integrand functions : $$\mathrm{the}\mathrm{first}:(_j)^2S_2\mathrm{exp}\{\mathrm{\hspace{0.17em}2}\frac{S_2}{\mathrm{}}\}=\frac{2}{\mathrm{}}(_jS_2)^2\mathrm{exp}\{\mathrm{\hspace{0.17em}2}\frac{S_2}{\mathrm{}}\}$$ (40) $$\mathrm{and}\mathrm{the}\mathrm{second}:(_j)^2S_2=\frac{2}{\mathrm{}}(_jS_2)^2$$ (41) Hence it is obviously seen that in a line with the existence of the eq. (41) the equation (33) could been rewritten in the following transparent form : $$\frac{S_1}{t}=\frac{1}{2m}(_jS_1)^2+\frac{1}{2m}(_jS_2)^2+U(r,t)$$ (42) In such a way it is evidently that the right-hand side expressions of the equations of the continuity (35) and (36) are canceled by the virtue of the eq:(41).Consequently we had an opportunity to shoe that the continuity partial differential equations are satisfied not only in the form (34), but they are satisfied also in the forms (35) and (36). Furthermore the expression of the eq.(42) might been interpreted from my new point of view, that the kinetic energy $`E_k`$ of the SchrEl is formed by two differential parts. Really, if the first part $`\frac{(_jS_1)^2}{2m}`$ describes the kinetic energy of its regular translation motion along some clear-cut thin smooth classical trajectory in an accordance with the laws of the NrlClsMch and ClsElcDnm with its current velocity $`V_j=\frac{1}{m}_jS_1`$ , then the second part $`\frac{(_jS_2)^2}{2m}`$ mouth describe the kinetic energy of its Furthian quantum stochastic motion of the FrthQntMicrPrt with its probable velocity $`U_j=\frac{1}{m}_jS_1`$ in a total analogy with the Brownian classical stochastic motion of the BrnClsMicrPrt with its osmotic velocity.Therefore it is very helpfully to rewrite the expression (42) in the following well-known form : $$E=\frac{mV^2}{2}+\frac{mU^2}{2}=\frac{(\overline{P})^2}{2m}+\frac{(\mathrm{\Delta }P)^2)}{2m}$$ (43) Indeed, some new facts have been brought to light. Therefore the upper investigation entitles us to make the explicit assertion that the most important difference between the quadratic differential wave equation in partial derivative of Schrodinger and the quadratic differential particle equation in partial derivative of Hamilton-Jacoby is exhibited by the existence of the kinetic energy of the QntMicrPrt’s Furthian trembling circular oscillations harmonic motion in the first one. $$\frac{S_1}{t}=\frac{(_jS_1)^2}{2m}+\frac{(_jS_2)^2}{2m}+U;$$ (44) As we can observe by cursory comparison there is a total coincidence of eq.(16) with eq.(44). Hence we are able to proof that the QdrDfrPrtEqt with PrtDrv of Schrodinger may be obtained from the QdrDfrPrtEqt with PrtDrv of Hamilton-Jacoby by addition the part of the kinetic energy of the Furthian stochastic circular harmonic oscillations motion. Indeed, it is obviously to understand that the first term $`\frac{(_lS_1)^2}{2m}`$ in the eq.(44) describes the kinetic energy of the regular translation motion of the NtnClsPrt with its current velocity $`V_l=\frac{_lS_1}{m}`$ and the second term $`\frac{(_lS_2)^2}{2m}`$ describes the kinetic energy of the random trembling circular harmonic oscillations motion (RndTrmMtn) of the FrthQntPrt in a total analogous with BrnClsPrt with its osmotic velocity $`U_l=\frac{_lS_2}{m}`$. Therefore we can rewrite the expression (44) in the following form : $$E_t=\frac{mV^2}{2}+\frac{mU^2}{2}+U=\frac{\overline{P}^2}{2m}+\frac{(\mathrm{\Delta }P)^2}{2m}+U;$$ (45) After elementary physical obviously suppositions some new facts have been brought to light. Therefore the upper investigation entitles us to make the explicit assertion that the most important difference between the QntQdrDfr WvEqt with PrtDrv of Schodinger and the ClsQdrDfrPrtEqt with PrtDrv of Hamilton-Jacoby is exhibited by the existence of the kinetic energy of the FrthRndTrmCrcHrmOscsMtn in the first one. Therefore when the SchEl is appointed in the Coulomb’s potential of the atomic nucleus spotted like (SptLk) elementary electric charge (ElmElcChrg) $`Ze`$ its total energy may be written in the following form : $$E_t=\frac{1}{2m}\left[(P_r)^2+\frac{(L)^2}{(r)^2}\right]+\frac{1}{2m}\left[(\mathrm{\Delta }P_r)^2+\frac{(\mathrm{\Delta }L)^2}{r^2}\right]\frac{Ze^2}{r}$$ (46) As any SchEl has eigenvalues $`n_r=\mathrm{\hspace{0.17em}0}`$ and $`l=\mathrm{\hspace{0.17em}0}`$ in a case of its ground state, so it follows that $`P_r=\mathrm{\hspace{0.17em}0}`$ and $`L=\mathrm{\hspace{0.17em}0}`$. As a consistency with the eq.(48) the eigenvalue of the SchEl’s total energy $`E_t^o`$ in its ground state in some H-like atom is contained only by two parts : $$E_t^o=\frac{1}{2m}\left[(\mathrm{\Delta }P_r)^2+\frac{(\mathrm{\Delta }L)^2}{(r)^2}\right]\frac{Ze^2}{r}$$ (47) Further the values of the dispersions $`(\mathrm{\Delta }P_r)^2`$ and $`(\mathrm{\Delta }L)^2`$ can be determined by virtue of the Heisenberg Uncertainty Relations (HsnUncRlt) : $$(\mathrm{\Delta }P_r)^2\times (\mathrm{\Delta }r)^2\frac{\mathrm{}^2}{4}$$ (48) $$(\mathrm{\Delta }L_x)^2\times (\mathrm{\Delta }L_y)^2\frac{\mathrm{}^2}{4}(\mathrm{\Delta }L_z)^2$$ (49) Thence the dispersion $`(\mathrm{\Delta }P_r)^2`$ will really have its minimal value at the maximal value of the $`(\mathrm{\Delta }r)^2==r^2`$.In this way the minimal dispersion value of the $`(\mathrm{\Delta }P_r)^2`$ can be determined by the following equation : $$(\mathrm{\Delta }P_r)^2=\frac{\mathrm{}^2}{4r^2}$$ (50) As the SchEl’s ground state has a spherical symmetry at $`l=\mathrm{\hspace{0.17em}0}`$,then the following equalities take place : $$(\mathrm{\Delta }L_x)^2=(\mathrm{\Delta }L_y)^2=(\mathrm{\Delta }L_z)^2;$$ (51) Hence we can obtain minimal values of the dispersions (51) through division of the eq.(48) with the corresponding equation from the eq. (51).In that a way we obtain the following result : $$(\mathrm{\Delta }L_x)^2+(\mathrm{\Delta }L_y)^2+(\mathrm{\Delta }L_z)^2=\frac{3\mathrm{}^2}{4}$$ (52) Just now we are in a position to rewrite the expression (48) in the handy form as it is well-known : $$E_t^o=\frac{1}{2m}\left[\frac{\mathrm{}^2}{4r^2}+\frac{3\mathrm{}^2}{4r^2}\right]\frac{Ze^2}{r}=\frac{1}{2}\frac{\mathrm{}^2}{mr^2}\frac{Ze^2}{r};$$ (53) It is extremely important to note here that we have used undisturbed ElcInt $`E_j`$ (3) of the QntElcMgnFld of StchVrtPhtns from the FlcVcm by dint of the equations (2) and (3) in order to obtain constrain of dynamical mutual conjugated quantities $`r_j`$ (7) and $`P_x`$ (8) from the NrlClsMch in their operator forms $`\widehat{r}_j`$ and $`\widehat{P}_j`$ within NrlQntMch. The quantum behaviour of the SchEl within NrlQntMch is caused by the ElcIntAct between its WllSpr ElmElcChrg and the ElcInt $`E_j`$ of the undisturbed QntElcMgnFld of StchVrtPhtns from the FlcVcm. So in consequence of the continuous ElcIntAct of the SchEl’s WllSpr ElmElcChrg with the ElcInt of the QntElcMgnFld of StchVrtPhtns one participates in the Furthian quantized stochastic motion (FrthQntStchMtn), which is quite obviously analogous of the Brownian classical stochastic motion (BrnClsStchMtn). As it is well-known the BrnClsPrts have no classical wave properties (ClsWvPrp), but the FrthQntPrts have QntWvPrps and display them every where. The cause of this distinction consists of indifference between the liquid and FlcVcm. Indeed, if atoms and molecules within liquid have no ClsWvPrps,all excitations of the FlcVcm and one itself have QntWvPrp. Therefore the FlcVcm transfers its QntWvPrp over the SchEl at ones ElcIntAct with its WllSpr ElmElcChrg.
warning/0001/astro-ph0001347.html
ar5iv
text
# Halo Profiles and the Nonlinear Two- and Three-Point Correlation Functions of Cosmological Mass Density ## 1 Introduction In this Letter we examine the power spectrum and the bispectrum for numerical simulations of the $`n=2`$ scale free model and cold dark matter (CDM) model. The power spectrum $`P(k)`$ is the most fundamental statistic for a random field and the only statistic for a Gaussian field. The cosmological distribution of matter, however, is not Gaussian, and higher order statistics are also important. The lowest-order non-Gaussian information is encoded in the bispectrum $`B(k_1,k_2,k_3)`$. In analogy to $`P(k)`$, which is the Fourier transform of the two-point correlation function $`\xi `$, the bispectrum is the Fourier transform of the three-point correlation function $`\zeta `$. The power spectrum $`P(k)`$ of matter fluctuations is related to the mass density field $`\delta =\rho /\overline{\rho }1`$ in $`k`$-space by $`\delta (\text{k}_1)\delta (\text{k}_2)=P(k_1)\delta _\mathrm{D}(\text{k}_1+\text{k}_2)`$, where $`\delta _\mathrm{D}`$ is the Dirac delta-function. The bispectrum $`B(k_1,k_2,k_3)`$, similarly, is related to $`\delta `$ by $`\delta (\text{k}_1)\delta (\text{k}_2)\delta (\text{k}_3)=B(k_1,k_2,k_3)\delta _\mathrm{D}(\text{k}_1+\text{k}_2+\text{k}_3)`$. On sufficiently large length scales, density fluctuations are small enough that perturbation theory for gravitational instability suffices, and the predictions for the correlation functions can be computed in detail. To linear order, the fluctuation amplitude $`\delta (\text{k},t)`$ grows by an overall scale factor, $`\delta =D(t)\delta _0(\text{k})`$, where $`\delta _0(\text{k})`$ is the amplitude at some early time $`t_0`$ and $`D(t)`$ is a growing function of time. In a nonlinear theory, interactions generate a sum of terms to all orders in the initial amplitude, $`\delta =\delta ^{(1)}+\delta ^{(2)}+\delta ^{(3)}+\mathrm{}`$, where $`\delta ^{(n)}\delta _0^n`$. Even for a Gaussian initial distribution, these nonlinear terms induce nonvanishing higher order correlations. In perturbation theory, the lowest order two-point statistic is the linear power spectrum $`P_l(k)[\delta ^{(1)}]^2`$. For the three-point statistic, the first nonvanishing contribution (i.e. tree-level) to the bispectrum is related to $`P_l`$ by $$B^{(0)}(k_1,k_2,k_3)=F(\text{k}_1,\text{k}_2)P_l(k_1)P_l(k_2)+F(\text{k}_2,\text{k}_3)P_l(k_2)P_l(k_3)+F(\text{k}_3,\text{k}_1)P_l(k_3)P_l(k_1),$$ (1) where $`F(\text{k}_i,\text{k}_j)=\frac{10}{7}+(k_i/k_j+k_j/k_i)(\widehat{\text{k}}_i\widehat{\text{k}}_j)+\frac{4}{7}(\widehat{\text{k}}_i\widehat{\text{k}}_j)^2`$ (Fry 1984). For $`\mathrm{\Omega }<1`$ the factors $`\frac{10}{7}`$ and $`\frac{4}{7}`$ become $`1\pm \kappa `$, where for $`\mathrm{\Omega }`$ near 1$`,\kappa \frac{3}{7}\mathrm{\Omega }^{2/63}`$ in an open universe and $`\kappa \frac{3}{7}\mathrm{\Omega }^{1/143}`$ in a flat universe with cosmological constant (Bouchet et al. 1995). Equation (1) indicates that the main dependence on the power spectrum can be removed if one defines the hierarchical three-point amplitude $$Q(k_1,k_2,k_3)\frac{B(k_1,k_2,k_3)}{P(k_1)P(k_2)+P(k_2)P(k_3)+P(k_3)P(k_1)}.$$ (2) The three-point amplitude $`Q`$ has the convenient feature that for the lowest nonvanishing result in perturbation theory (eq. ), $`Q^{(0)}`$ is independent of time and the overall amplitude of $`P_l`$. Moreover, for scale-free models with a power-law $`P_l(k)`$, $`Q^{(0)}`$ is independent of overall scale as well. In a pure hierarchical model, $`Q`$ is exactly constant. The bispectrum $`B`$ and the three-point amplitude $`Q`$ depend on any three parameters that define a triangle in $`k`$-space. For simplicity, we will use equilateral triangles to explore scale dependence. We have examined several other configurations and found behavior similar to the equilateral case. For a closed equilateral triangle, one has $`k_1=k_2=k_3=k`$ and $`\widehat{\text{k}}_i\widehat{\text{k}}_j=\frac{1}{2}`$, and the bispectrum $`B_{\mathrm{eq}}`$ depends only on the single wavenumber $`k`$. To lowest order, the equilateral bispectrum has a particularly simple form, $`B_{\mathrm{eq}}^{(0)}(k)=\frac{12}{7}P_l^2(k)`$, and it follows from equation (2) that $`Q_{\mathrm{eq}}^{(0)}(k)=\frac{4}{7}`$, independent of the power spectrum. Since $`B`$ has dimensions of $`\mathrm{Mpc}^6`$, it is convenient to define a new dimensionless variable, which we name $`\mathrm{\Delta }_3`$, to characterize the three-point statistic of mass density fluctuations. Following the familiar density variance $`\mathrm{\Delta }(k)=4\pi k^3P(k)`$ for the two-point statistic, we define $`\mathrm{\Delta }_3`$ for equilateral triangles as $$\mathrm{\Delta }_3(k)4\pi k^3\sqrt{B_{\mathrm{eq}}(k)};$$ (3) and it follows that $`\mathrm{\Delta }_3(k)/\mathrm{\Delta }(k)=\sqrt{3Q_{\mathrm{eq}}(k)}`$. Since $`Q_{\mathrm{eq}}`$ is simply a constant to the lowest nonvanishing order in perturbation theory, this new statistic also has the convenient property that the leading order $`\mathrm{\Delta }_3^{(0)}(k)`$ and $`\mathrm{\Delta }^{(0)}(k)`$ have identical shapes, i.e., $`\mathrm{\Delta }_3^{(0)}(k)/\mathrm{\Delta }^{(0)}(k)=(12/7)^{1/2}`$. ## 2 High-$`k`$ Behavior The leading-order perturbative expressions given above are valid for $`\delta 1`$, or at low $`k`$. Including the next-to-leading order terms for the bispectrum extends the range of validity to the quasilinear regime $`\delta 1`$ (Scoccimarro et al. 1998). Numerical simulations, however, are required in order to determine the fully evolved bispectrum of mass density fluctuations into the deeply nonlinear regime at high $`k`$. We examine results from two cosmological simulations, an $`n=2`$ scale-free model and a cold dark matter (CDM) model. The $`n=2`$ simulation has $`256^3`$ particles and a Plummer force softening length of $`L/5120`$, where $`L`$ is the box length. This is the highest resolution simulation of an $`n=2`$ model to our knowledge. It is also the main simulation used in the study of self-similar scaling of the power spectrum by Jain & Bertschinger (1998). The CDM model is critically flat with matter density $`\mathrm{\Omega }_m=0.3`$ and cosmological constant $`\mathrm{\Omega }_\mathrm{\Lambda }=0.7`$. This run has $`128^3`$ particles and is performed in a $`(100\mathrm{Mpc})^3`$ comoving box with a comoving force softening length of $`50\mathrm{kpc}`$ for Hubble parameter $`h=0.75`$. The baryon fraction is set to zero for simplicity. The primordial power spectrum has a spectral index of $`n=1`$, and the density fluctuations are drawn from a random Gaussian distribution. The gravitational forces are computed with a particle-particle particle-mesh (P<sup>3</sup>M) code (Ferrell & Bertschinger 1994). We compute the density field $`\delta `$ on a grid from particle positions using the second-order triangular-shaped cloud (TSC) interpolation scheme. A fast Fourier transform is then used to obtain $`\delta `$ in $`k`$-space. The $`k`$-space TSC window function is deconvolved to correct for smearing in real space due to the interpolation, and shot noise terms are subtracted to correct for discreteness effects. Figure 1 illustrates the behavior of the nonlinear power spectrum $`\mathrm{\Delta }(k)`$ and the three-point statistics $`\mathrm{\Delta }_3(k)`$ and $`Q(k)`$ for equilateral triangular configurations for the $`n=2`$ scale-free simulation. Three time outputs are shown, and the results are plotted against scaled $`k/k_{nl}`$, where $`k_{nl}`$ is defined by $`_0^{k_{nl}}d^3kP_l(a,k)=1`$, and the expansion factor $`a`$ is 1 initially. Unlike the two-point $`\mathrm{\Delta }`$, for which self-similar scaling works well (Jain & Bertschinger 1998), the three-point curves in the plot do not overlap; self-similarity therefore does not hold as rigorously for the three-point function in the simulations. The slope of $`Q`$ also has a weak dependence on time. For the last output ($`a=26.91`$, $`k_{nl}=7.25`$), $`Q`$ increases monotonically with $`k`$ down to the resolution limit with an approximate logarithmic slope of $`0.225`$. The next-to-leading order perturbative results for $`Q`$ (Scoccimarro et al. 1998) roughly agree with the most weakly evolved, ($`a=13.45`$, $`k_{nl}=29`$) output up to $`kk_{nl}`$, but not at higher $`k`$ nor in the later outputs. The upturn of $`Q_{\mathrm{eq}}`$ at high $`k`$ (dotted portion of curves in Fig. 1) is related to the downturn in $`\mathrm{\Delta }`$ at the same $`k`$, which is likely due to limited numerical resolution. Many earlier papers did not report this behavior (e.g., Efstathiou et al. 1988; Fry, Melott, & Shandarin 1993; Scoccimarro et al. 1998). Some of the computations were not carried out to sufficiently high wavenumbers to provide a clear test of the hierarchical model. Others (e.g. Scoccimarro et al. 1998) assumed self-similarity and averaged results over different time outputs, treating each output as if it were from an independent realization. We have checked for possible numerical artifacts by re-examining the $`n=2`$ results of Fry et al. (1993) from $`128^3`$ particle-mesh (PM) simulations. At evolution stages comparable to those in Figure 1 (i.e. $`k_{nl}=32`$, 16, and 8), we find that $`Q_{\mathrm{eq}}`$ in four different random realizations agrees well with Figure 1 up to the resolution limit of the PM runs. Departures from the hierarchical model have been suggested in several papers that have examined the real-space three-point function (Suto & Matsubara 1994) and the skewness $`S_3`$ (e.g., Lahav et al. 1993; Colombi et al. 1996), a volume integral average of the full three-point function. The simulations used in these papers had $`64^3`$ particles and a smaller dynamic range, so the results were limited to lower $`k`$. The skewness is also a less powerful probe of hierarchical clustering than the full three-point examined here because the volume integral might average out non-hierarchical signals. ## 3 Dependence on Halo Density Profiles In this section we examine analytical models that can be used to elucidate the numerical results of Figure 1. We suggest that the key to a better understanding of $`Q`$ in the high-$`k`$ regime lies in the connection between $`Q`$ and the density profiles of dark matter halos. This is because the contributions to $`Q`$ at high $`k`$ are dominated by close particle triplets, many of which are located within individual halos of size $`2\pi /k`$. It is therefore reasonable to expect the two- and three-point correlation functions on small length scales to reflect the density profiles of dark matter halos. Sheth and Jain (1997) have examined how the amplitude of the two-point function is related to simple halo profiles. Our emphasis here is on the link between the shape of the three-point function and halo profiles. One simple way to demonstrate the relation between halos and the correlation functions is to consider the “power law cluster” model. In this model, matter is distributed in randomly placed halos with a power-law mass density profile $`\rho (r)r^ϵ`$ out to some radius $`R`$ (Peebles 1974; 1980). For $`1.5<ϵ<3`$ and $`r<R`$, one can derive analytically that the two-point function $`\xi `$, the density variance $`\mathrm{\Delta }`$, the three-point function $`\zeta `$, and $`Q`$ scale as $$\xi (r)r^{32ϵ},\mathrm{\Delta }(k)k^{2ϵ3},\zeta (r)r^{33ϵ},Q\frac{\zeta }{\xi ^2}r^{ϵ3}k^{3ϵ}.$$ (4) This model is clearly simplistic because in realistic cosmological models, the centers of dark matter halos are not randomly distributed, the halo density profile is not a pure power law, and not all mass is located within the halos. Equation (4) nonetheless provides valuable insight into the link between the shape of dark matter halos and the high-$`k`$ behavior of the two- and three-point clustering statistics. To test the applicability of equation (4) in a more realistic setting, we have performed a series of experiments with dark matter halos identified in cosmological $`N`$-body simulations. In these experiments, we keep the centers and masses of the halos unchanged but redistribute the subset of particles which lies within the virial radius (the radius within which $`\delta >200`$) of each halo according to $`\rho r^ϵ`$ for a chosen $`ϵ`$. We then recompute $`\mathrm{\Delta }`$ and $`Q_{\mathrm{eq}}`$ from the particle positions. This recipe allows us to explore simple density profiles while preserving the actual halo-halo spatial correlations, the halo mass distribution, and the correlation signal from non-halo particles in the simulations. Figure 2 compares the results for the original output in the $`n=2`$ scale-free simulation with the results for replaced power-law halos. Figure 3 does the same for the $`\mathrm{\Omega }_m=0.3`$ CDM simulation. As indicated, when the halo profiles obey $`\rho r^ϵ`$, we find that $`\mathrm{\Delta }`$ and $`Q`$ at high $`k`$ are well approximated by power laws, and that the slopes are indeed related to $`ϵ`$ according to equation (4). For $`ϵ=2`$ (i.e. an isothermal density distribution), for example, both Figures 2 and 3 show that $`\mathrm{\Delta }k`$ and $`Q_{\mathrm{eq}}k`$. Figure 3 also verifies equation (4) for $`ϵ=2.35`$: $`\mathrm{\Delta }k^{1.7}`$ and $`Q_{\mathrm{eq}}k^{0.65}`$. We therefore conclude that for a power-law density profile, equation (4) works well at predicting the slopes of the power spectrum and the three-point $`Q`$ in the nonlinear regime. This remains true even when the clustering of the halo centers and the correlation signal from non-halo particles are taken into account. ## 4 Discussion To gain a better understanding of the behavior of two- and three-point statistics in the deeply nonlinear regime, we have investigated in this Letter how the shapes of dark matter halos affect the power spectrum $`P`$ and the three-point amplitude $`Q`$ at high $`k`$. For halos with power-law density profiles, Figures 2 and 3 demonstrate that the simple model of equation (4) works very well at predicting the slopes of $`P`$ and $`Q`$ at high $`k`$. The figures also show that the converse is not true: the slope of $`P`$ or $`Q`$ alone does not determine a unique halo shape. For example, Figure 2 shows that $`P`$ at high $`k`$ in the $`n=2`$ scale free model is only slightly changed when simulation halos are replaced with $`ϵ=2`$ power-law halos. It would therefore be difficult to distinguish these two profiles using the power spectrum alone. The three-point $`Q`$, however, is drastically different for the two profiles. This indicates that although the nonlinear $`P`$ and $`Q`$ are both closely related to the shapes of halos, they probe different regions and properties of the halos. A plausible explanation for this difference is that the two-point function reflects the mean density at a given distance from a particle, whereas the three-point function reflects the dispersion in density at the given scale (Peebles 1980, § 36). Realistic dark matter halos do not have pure power law profiles. High-resolution $`N`$-body simulations have shown that cold dark matter halos have a roughly (Jing & Suto 2000) universal density profile with $`ϵ=3`$ at the outer radii and $`ϵ=1`$ (Navarro, Frenk & White 1997; Huss, Jain, & Steinmetz 1999) or $`ϵ=1.5`$ (Moore et al. 1999) nearer the halo centers. A further complication is that these profiles have a mass-dependent “concentration parameter”: less massive halos are more centrally concentrated. It will be useful to develop a model more extensive than equation (4) to incorporate these factors. Such a model should help to shed light on why $`Q(k)`$ at high $`k`$ in our figures is approximately a power law, even though the halo profile in the simulations is not a pure power law. Given that the shape of $`Q`$ is highly sensitive to halo profiles, the model can also be used to tackle the question: what criteria must the halos satisfy in order for hierarchical clustering to occur? In the best simulation of the $`n=2`$ scale-free model available to us (with 16.7 million particles and a force resolution of $`L/5120`$), we find that the three-point amplitude $`Q`$ is not a constant with wavenumber $`k`$ and that the self-similar scaling observed for the two-point function (Jain & Bertshinger 1998) does not hold as rigorously for the three-point function (Fig. 1). We have checked for possible numerical artifacts using smaller simulations with different phases (see § 2), and the results are similar within the range of overlap but may still be a manifestation of the finite simulation volume, to which the $`n=2`$ spectrum is particularly susceptible. Based on lower resolution simulations for scale-free initial conditions with spectral index $`n`$, Fry et al. (1993) suggested $`Q=3/(3+n)`$ as the asymptotic value in the nonlinear regime. Scoccimarro & Frieman (1999) in so-called hyperextended perturbation theory obtain $`Q=(42^n)/(1+2^{n+1})`$. These two expressions would predict $`Q=3`$ and 2.5, respectively, for $`n=2`$ at high $`k`$. Figure 1 shows that $`Q`$ reaches a similar value between 2.5 and 3 at $`kk_{nl}`$, but it rises to 4 to 5 at $`k10k_{nl}`$. For CDM-type models, the effective index decreases with $`k`$, so the two expressions above would give $`Q`$ rising with $`k`$, but in neither case is the prediction as strong as what we see. It is useful to compare our numerical results with theoretical expectations. From the assumptions of stable clustering and self-similarity, it is well known that for the two-point function, $`\mathrm{\Delta }(k)k^{(9+3n)/(5+n)}`$ at high $`k`$ for an initial $`P_lk^n`$ (Davis & Peebles 1977). The nonlinear $`\mathrm{\Delta }`$ in Figure 1 and earlier work of Efstathiou et al. (1988), Peacock & Dodds (1996), Jain (1997), Ma (1998), and Jain & Bertschinger (1998) all support this behavior. For the three-point function, these same conditions would result in constant $`Q`$, i.e., hierarchical clustering. The fact that high-resolution simulation results do not support this implies that one of these assumptions – stability or self-similarity – must be violated, or still better future simulations are needed. ###### Acknowledgements. We are grateful to John Peacock for enlightening discussions and Edmund Bertschinger for providing the $`n=2`$ scale-free simulation. Computing time for this work is provided by the National Scalable Cluster Project and the Intel Eniac2000 Project at the University of Pennsylvania. C.-P. M. acknowledges support of an Alfred P. Sloan Foundation Fellowship, a Cottrell Scholars Award from the Research Corporation, a Penn Research Foundation Award, and NSF grant AST 9973461.
warning/0001/astro-ph0001022.html
ar5iv
text
# A multipole-Taylor expansion for the potential of gravitational lens MG J0414+0534 ## 1 Introduction There are now about 40 observed examples of “strong” gravitational lensing, in which multiple images of a background object are produced by the gravitational potential of an intervening object (Kochanek et al. (1999)). Constructing models for the gravitational potential of the deflecting mass in these systems is of fundamental importance, in the first place to verify that the observed system is truly a gravitational lens. A crucial hurdle for every promising gravitational lens candidate is to have its image configuration reproduced, at least approximately, by a plausible model for the deflecting mass distribution. If this cannot be achieved, the lensing interpretation must be seriously doubted. Once this and other tests for lensing are passed successfully, however, model construction moves beyond a plausibility check into a direct measurement of the mass distribution of the lens. Lensing makes a unique contribution because it is sensitive to all forms of mass (including dark matter), yet does not depend on any luminous tracers in the lens. Lens modeling is also a crucial step in the enterprise of determining cosmological parameters by measuring the time delays between the light curves of multiple images. A successful model will predict the values of these time delays in terms of parameters such as $`H_0`$, $`\mathrm{\Omega }_m`$, and $`\mathrm{\Omega }_\mathrm{\Lambda }`$. These parameters can then be constrained by the time-delay measurements (Refsdal 1964, (1966); for recent examples see Haarsma et al. (1999), Lovell et al. (1998)). The appeal of this technique is that it does not make use of the usual chain of intermediate distance indicators and their associated uncertainties. The determinations of mass distributions and cosmological parameters are both frustrated by the main challenge of lens modeling: it is a poorly-constrained inverse problem with unknown systematic errors. It is neither obvious which parameterized model for the gravitational potential should be chosen, nor how precisely the parameters are constrained by the data. The choice of parameterization for the gravitational potential is usually dictated by two factors: preconceptions about the mass distribution of galaxies (based on the distribution of luminous matter and/or velocity dispersions), and ease of computation. Some examples are singular isothermal spheres, elliptical density profiles, elliptical potential contours, and triaxial density profiles (see e.g. Schneider & Weiss (1991); Nair & Garrett (1997); Chae, Khersonksy & Turnshek (1998)). Added to these are terms representing external “shear” from the tidal forces of neighboring mass concentrations. Such models have all been used successfully to reproduce the image configurations of the known sample of lenses, at least qualitatively, which leads one to wonder how well the observations of strong lensing actually constrain the lens potential. Kochanek (1991) was the first to investigate this question systematically. He attempted to model each of 12 different lenses using a set of 6 different parameterized potentials. The potentials he considered were point masses and singular isothermal spheres, added to shear terms with one of three radial dependencies. He found that the total mass in the region between the multiple images is well constrained, but the radial dependence of the potential is not. This was because the multiple images are typically located close to the “Einstein ring” of the lens, so the constraints on the potential are correspondingly limited to a small range in radius. Kochanek suggested (but did not carry out) a Taylor expansion, parameterized by the deviation from the ring radius, as a way to turn this fact to advantage. The purpose of this paper is to elaborate upon this suggestion and apply it to a particular quadruple-image gravitational lens, MG J0414+0534. We expand the potential as a series of multipoles in angle, and as a modified Taylor series in radius. The resulting series has three advantages over traditional lens models. One, it is mathematically general, and therefore less subject to preconceptions about the galactic mass distribution (which, after all, is what is trying to be determined). Two, the degeneracies in the lens equation and the moments of the mass distribution have simple correspondences to individual terms of the series. Three, each successive term in the series is expected to contribute a smaller amount to the light deflection, so long as the radial parameter is small and the angular dependence of the potential is smooth. This allows the complexity of the model to be easily prescribed by the truncation of the series. However, although the first assumption (the smallness of the radial parameter) is empirically true for quadruple-image lenses, the second assumption (the angular smoothness of the potential) is open to question. Nevertheless, in the face of our ignorance of true galaxy structure, we employ a multipole expansion because of its simplicity and generality. The disadvantage of a mathematically general expansion (rather than one that is motivated by astrophysical preconceptions) is that the number of parameters is large. In many cases, especially for the double-image lenses, the number of parameters would be comparable to or more than the number of observational constraints. This is probably why such an expansion has not been used previously. However, recent VLBI images of MG J0414+0534 provide a much larger body of constraints than were previously available (Trotter (1998)). Each of the four lensed images was found to contain four components, with clear correspondences between the components of different images. This provides 16 components whose positions are known with milliarcsecond precision, thereby creating a testbed for the multipole-Taylor expansion. This paper will be organized as follows. The next section describes prior optical and radio observations of MG J0414+0534, in addition to the latest VLBI map. Section 3 presents the formalism for our series expansion, discusses the physical significance of each term, and identifies the terms that cannot be constrained due to degeneracies in the lens equation. Some previous models for MG J0414+0534 are discussed and compared to our technique. Section 4 explains the numerical methods we employed and presents the best-fit results. Finally, section 5 discusses the implications of these results for the mass distribution of the lens and the time delays between images. ## 2 Observations of MG J0414+0534 The radio source MG J0414+0534 was first identified by Hewitt et al. (1992) as a gravitational lens during a systematic search for lenses in the MIT-Green Bank 5 GHz radio catalog. In both optical and radio images it has four components, which have been called A1, A2, B, and C, in order of decreasing brightness. The radio components all have spectral index<sup>1</sup><sup>1</sup>1The spectral index $`\alpha `$ is defined such that $`S_\nu \nu ^\alpha `$, where $`S_\nu `$ is the spectral flux density. $`\alpha =0.80\pm 0.02`$ (Katz, Moore & Hewitt (1997)), and the optical components are all exceedingly red. However, the A1/A2 radio flux ratio ($`1.1`$) and optical flux ratio ($`2.5`$) do not agree, even though gravitational light deflection is independent of wavelength. The discrepancy could be caused by a number of factors (Angonin-Willaime et al. (1994)), including dust (McLeod et al. (1998)) and microlensing (Witt, Mao & Schechter (1995)). Schechter & Moore (1993) discovered the lensing galaxy in the $`I`$ band, along with a faint object $`1\mathrm{}`$ west of component B which they named object X. The optical/infrared spectrum of the lensed source resembles a very reddened quasar at $`z_s=2.639\pm 0.002`$ (Lawrence et al. (1995)). The redshift of the lensing galaxy is $`z_l=0.9584\pm 0.0002`$ (Tonry & Kochanek (1998)). The best presently published optical photometry (uncertainty $`0.004`$ mag) and astrometry (uncertainty $`0.02\mathrm{}`$) for this system were obtained with Hubble Space Telescope (HST) WFPC2/PC1 observations by Falco, Lehár & Shapiro (1997), who also detected a blue arc extending from A1 to A2 to B. The deepest radio observations to date did not detect a fifth radio component in the system (Katz, Moore & Hewitt (1997)). Higher-resolution radio maps, using MERLIN (Garrett et al. (1993)) and VLBI (Patnaik & Porcas (1996)), showed substructure within the four images of MG J0414+0534. In particular, the VLBI map resolved images A1, A2, and B into two components each, and revealed image C to be extended. Previous attempts to use VLBI constraints to perform lens modeling, however, were hampered by the low signal-to-noise ratio of the available data (Ellithorpe (1995)). As part of a 1995 study to test the feasibility of using the NRAO<sup>2</sup><sup>2</sup>2The National Radio Astronomy Observatory (NRAO) is operated by Associated Universities, Inc., under cooperative agreement with the National Science Foundation. Very Long Baseline Array (VLBA) to monitor the time-variability of various radio-loud lenses, we obtained 5 GHz images of MG J0414+0534. These observations and the data reduction procedure are described in detail elsewhere (Trotter (1998)) and will be presented in a future paper; only the results are summarized here. The synthesized beam was $`1.5\times 3.5`$ milliarcseconds, and the RMS thermal noise was about 0.15 mJy/beam, close to the theoretical limit. The peak fluxes in the maps of each of the four images ranged from 21 to 110 mJy/beam, allowing a much higher signal-to-noise ratio than was previously available. Within each image were detected four components. The compact components p, q, and r were labeled in order of decreasing brightness; there is also one extended component, labeled s. The four maps are shown in Figure 1, superimposed on a lower-resolution VLA map. The sub-components are labeled in the larger maps shown in Figure 2. The locations, fluxes, and extents of the components were determined by least-squares fitting to a set of elliptical Gaussian parameters. Table 1 lists the locations and fluxes of the 16 VLBI components. ## 3 The modified multipole-Taylor expansion The motivation behind our method for lens modeling is to expand the lens potential in a manner which preserves a measure of mathematical generality and in which the terms that can and cannot be constrained by lensing data are clearly separated. We have tried to arrange for the parameters in our expansion to be sensitive to what can be learned from lensing data, rather than what one would wish a priori to learn. In this section we will develop this expansion in detail, describe the physical significance and constrainability of each term, and compare our expansion to other commonly-used parameterizations. ### 3.1 The multipole-Taylor expansion The goal of lens modeling is to deduce the two-dimensional gravitational lens potential, which is defined in a way similar to the definitions of Schneider, Ehlers & Falco (1992) and Narayan & Bartelmann (1996): $$\mathrm{\Phi }(𝐫)=\frac{2D_{LS}}{D_LD_S}𝑑z\mathrm{\Phi }_N(D_L𝐫,z).$$ (1) Here $`𝐫`$ is the angular coordinate measured from an arbitrary optic axis, $`z`$ is the line-of-sight coordinate, and $`\mathrm{\Phi }_N`$ is the Newtonian gravitational potential of the lens. The angular diameter distances $`D_L`$, $`D_S`$, and $`D_{LS}`$ are from observer to lens, observer to source, and lens to source, respectively. As elsewhere in this paper, the speed of light has been set to 1 by a choice of units. The lens potential is related to the surface mass density of the lensing object by the two-dimensional Poisson equation, $$_𝐫^2\mathrm{\Phi }(𝐫)=8\pi G\frac{D_LD_{LS}}{D_S}\sigma (𝐫)=\frac{2\sigma (𝐫)}{\sigma _\mathrm{c}}.$$ (2) Implicit in the above equation is the definition of the critical surface mass density, $`\sigma _\mathrm{c}=(1/4\pi G)(D_S/D_LD_{LS})`$. The lens potential determines the image configuration via the lens equation, $$𝐬=𝐫\mathbf{}_𝐫\mathrm{\Phi }(𝐫),$$ (3) where $`𝐫`$ is the image position and $`𝐬`$ is the source position. The typical procedure in lens modeling is to adopt a parameterized form for $`\mathrm{\Phi }`$, and then fix the parameters by minimizing an error function $`\chi ^2`$. The error function may represent the deviations between the observed images and the images that are projected through the model potential from a model source distribution. Alternatively, the error function may represent the deviations between source positions that are back-projected from the various images (which is computationally simpler), as explained further in section 55. The parameterization we adopt is a multipole-Taylor expansion. The first step is to expand the potential $`\mathrm{\Phi }(𝐫)`$ in terms of a complete set of orthogonal basis functions which make the Poisson equation separable: $$\mathrm{\Phi }(r,\theta )=\mathrm{\Phi }_0(r)+\underset{m=1}{\overset{\mathrm{}}{}}𝚽_m(r)(\widehat{𝐱}\mathrm{cos}m\theta +\widehat{𝐲}\mathrm{sin}m\theta )$$ (4) The function $`\mathrm{\Phi }_0(r)`$ is the monopole, and the vector-like functions $`𝚽_m(r)`$ are the higher multipoles. Schneider & Weiss (1991) also carried out a multipole expansion of the lens potential in this manner. The motivation for this expansion is that each multipole moment of the potential depends only on the corresponding multipole moment of the surface mass density. Specifically, the relations are (Kochanek (1991)): $`\mathrm{\Phi }_0(r)`$ $`=`$ $`\mathrm{constant}+2\mathrm{ln}r{\displaystyle _0^r}𝑑r^{}r^{}{\displaystyle \frac{\sigma _0(r^{})}{\sigma _\mathrm{c}}}+2{\displaystyle _r^{\mathrm{}}}𝑑r^{}r^{}\mathrm{ln}r^{}{\displaystyle \frac{\sigma _0(r^{})}{\sigma _\mathrm{c}}}`$ (5) $`𝚽_m(r)`$ $`=`$ $`{\displaystyle \frac{r^m}{m}}{\displaystyle _0^r}𝑑r^{}r^{(1+m)}{\displaystyle \frac{𝝈_m(r^{})}{\sigma _\mathrm{c}}}{\displaystyle \frac{r^m}{m}}{\displaystyle _r^{\mathrm{}}}𝑑r^{}r^{(1m)}{\displaystyle \frac{𝝈_m(r^{})}{\sigma _\mathrm{c}}}.`$ (6) For smooth mass distributions, only the few lowest-order terms in the multipole expansion of the surface mass density are expected to be significant, and the rest may be neglected. By reference to equations 5 and 6 it is clear that this is equivalent to neglecting the higher multipole components of the potential. For modeling purposes, one may truncate the expansion of the potential at the desired level. The radial dependence of $`\mathrm{\Phi }_0(r)`$ and $`𝚽_m(r)`$ must also be parameterized to be suitable for use in lens modeling. Lensed images constrain $`\mathbf{}\mathrm{\Phi }`$ and $`_i_j\mathrm{\Phi }`$ at the locations of the images. For Einstein rings and quadruple-image lenses (“quads”), the images are typically near the lens’s characteristic ring radius. Therefore, following the suggestion of Kochanek (1991), we expand the radial dependence of each multipole moment of the potential as a Taylor series in the parameter $`\rho =(rb)/b`$, where $`b`$ is the Einstein ring radius. (The meaning of the Einstein ring radius for a non-circular lens is discussed below, in section 3.2.) This series should converge quickly for image positions located at $`rb`$. For example, $`|\rho |<0.2`$ for all of the components in MG J0414+0534. The resulting expansion of the lens potential is: $`{\displaystyle \frac{1}{b^2}}\mathrm{\Phi }(r,\theta )`$ $`=`$ $`\mathrm{const}+\rho +{\displaystyle \frac{1}{2}}\rho ^2f_2+{\displaystyle \underset{t=3}{\overset{\mathrm{}}{}}}{\displaystyle \frac{1}{t!}}\rho ^tf_t`$ $`{\displaystyle \underset{m=1}{\overset{\mathrm{}}{}}}\left(\left\{𝐌_m^{\mathrm{sum}}+\rho m𝐌_m^{\mathrm{diff}}+{\displaystyle \underset{t=2}{\overset{\mathrm{}}{}}}{\displaystyle \frac{1}{t!}}\rho ^t𝐅_{mt}\right\}(\widehat{𝐱}\mathrm{cos}m\theta +\widehat{𝐲}\mathrm{sin}m\theta )\right),`$ parametrized by the origin of the expansion $`(g_x,g_y)`$, the ring radius $`b`$, the monopole parameters $`f_t`$, and the higher multipole parameters $`𝐌_m^{\mathrm{sum}}`$, $`𝐌_m^{\mathrm{diff}}`$, and $`𝐅_{mt}`$. We use $`t`$ as an integral index, contrary to convention, because of the mnemonic value of associating $`t`$ with Taylor and $`m`$ with multipole. The additive constant $`f_0`$ does not affect light deflection. Because we factor out the ring radius $`b`$, and then use it as a parameter, we have $`f_1=1`$. The $`m1`$ multipoles with $`t=0`$ and $`t=1`$ received the special names $`𝐌_m^{\mathrm{sum}}`$ and $`𝐌_m^{\mathrm{diff}}`$ for reasons that will become apparent in the next section. The model parameters $`g_x`$ and $`g_y`$, which specify the location of the origin of coordinates, are implicit in equation 3.1. To simplify the interpretation of the multipole moments, this origin should be centered on the deflector. In section 3.4 it will be shown how this condition may be enforced during the model-fitting procedure. ### 3.2 Physical significance of the expansion parameters Although the potential $`\mathrm{\Phi }(𝐫)`$ is the quantity most directly constrained by observations of lensing, it is the surface mass density $`\sigma (𝐫)`$ that is usually of direct astrophysical interest. In this section the correspondence between the the parameters in the multipole-Taylor expansion of $`\mathrm{\Phi }(𝐫)`$ and the multipoles of $`\sigma (𝐫)`$ will be made explicit. This is useful because the multipole moments of the surface mass density have a simple physical meaning. This meaning will be reviewed first, then the correspondence between the multipole-Taylor parameters and the surface mass density multipoles will be given. When $`\sigma (𝐫)`$ is expanded in a multipole series as in equation 4, it can be shown that: $`\sigma _0(r)`$ $`=`$ $`{\displaystyle \frac{1}{2\pi }}{\displaystyle _0^{2\pi }}𝑑\theta \sigma (r,\theta )`$ (8) $`𝝈_m(r)`$ $`=`$ $`{\displaystyle \frac{1}{\pi }}{\displaystyle _0^{2\pi }}𝑑\theta \sigma (r,\theta )(\widehat{𝐱}\mathrm{cos}m\theta +\widehat{𝐲}\mathrm{sin}m\theta ).`$ (9) These expressions make the physical meaning of the multipoles clear. The monopole moment $`\sigma _0(r)`$ is the average surface mass density in an infinitesimally narrow annulus of radius $`r`$, or, equivalently, the angularly-averaged surface mass density. The multipoles $`𝝈_m(r)`$ describe the distribution of matter around that annulus. In particular, the dipole moment $`𝝈_1(r)`$ points to the center of mass $`𝐱_c`$ of the annulus, viz., $$\frac{𝐱_c}{r}=\frac{1}{2}\frac{𝝈_1(r)}{\sigma _0(r)}.$$ (10) Likewise, the quadrupole moment $`𝝈_2(r)`$ arises from an elongated mass distribution. The $`𝝈_3(r)`$ term arises from any triangularity in the mass distribution. Any quadrangularity, e.g. boxiness or diskiness, will give rise to a nonzero $`𝝈_4(r)`$ term. The multipole $`𝝈_m(r)`$ can also be expressed as a magnitude $`|𝝈_m(r)|`$ and an angle $`\psi _{\sigma _m}(r)`$, $$𝝈_m(r)=|𝝈_m(r)|(\widehat{𝐱}\mathrm{cos}m\psi _{\sigma _m}(r)+\widehat{𝐲}\mathrm{sin}m\psi _{\sigma _m}(r)),$$ (11) rather than by its $`\widehat{𝐱}`$ and $`\widehat{𝐲}`$ components. Twisted isodensity contours cause a change in the angle $`\psi _{\sigma _m}(r)`$ with radius. In particular, a relative excess of mass at radius $`r`$ in any of the $`m`$ directions $`\psi _{\sigma _m}(r)+2\pi n/m`$ (where $`nm`$ is a positive integer) will contribute to the multipole moment $`𝝈_m(r)`$. It follows from the positivity of the surface mass density that $`|𝝈_m(r)|2\sigma _0(r)`$. The multipole moment attains its maximum amplitude only if all the mass in the annulus at radius $`r`$ is clumped at the equally spaced angles $`\psi _{\sigma _m}(r)+2\pi n/m`$, though it need not be evenly divided between these angles. In this manner, knowledge of the mass density’s multipole moments gives a direct picture of the location of the mass. We now relate the expansion parameters of the lens potential to the multipole moments of the surface mass density, by expanding equations 5 and 6 in a Taylor series about $`r=b`$. For the monopole, $$f_t\frac{1}{b^2}b^t\frac{d^t}{dr^t}\mathrm{\Phi }_0(r)|_{r=b},$$ (12) of which the first four terms are: $`f_0{\displaystyle \frac{1}{b^2}}\mathrm{\Phi }_0(b)`$ $`=`$ $`\mathrm{const}`$ (13) $`f_1{\displaystyle \frac{1}{b^2}}b{\displaystyle \frac{d}{dr}}\mathrm{\Phi }_0(r)|_{r=b}`$ $`=`$ $`1`$ (14) $`f_2{\displaystyle \frac{1}{b^2}}b^2{\displaystyle \frac{d^2}{dr^2}}\mathrm{\Phi }_0(r)|_{r=b}`$ $`=`$ $`1+{\displaystyle \frac{2\sigma _0(b)}{\sigma _\mathrm{c}}}`$ (15) $`f_3{\displaystyle \frac{1}{b^2}}b^3{\displaystyle \frac{d^3}{dr^3}}\mathrm{\Phi }_0(r)|_{r=b}`$ $`=`$ $`2{\displaystyle \frac{2\sigma _0(b)}{\sigma _\mathrm{c}}}+2b{\displaystyle \frac{d}{dr}}\left({\displaystyle \frac{\sigma _0(r)}{\sigma _\mathrm{c}}}\right)|_{r=b}`$ (16) As mentioned previously, the constant term $`f_0`$ can be ignored because a constant offset in the potential does not affect light deflection, nor does it affect the differential time delay between two images of the same source. (Although $`f_0`$ does affect the propagation time of the light from source to observer, only the relative or differential delay between two different images can be measured.) The linear term, $`f_1`$, sets the ring radius $`b`$ and does not have any other adjustable coefficient. In fact, equation 14 defines the Einstein ring radius $`b`$ not only for this expansion, but also for a general potential. It can be shown that the average surface density within the radius $`b`$ so defined is the critical surface density $`\sigma _\mathrm{c}`$ (Schneider, Ehlers & Falco (1992)). The quadratic term $`f_2`$ gives the angularly-averaged surface mass density at the ring radius. In the next section we will show that this term cannot be constrained by lensing phenomena, due to the mass-sheet degeneracy. The cubic term $`f_3`$ is therefore the first constrainable term yielding information about the radial profile of the mass distribution. The correspondence between the higher multipoles of $`\mathrm{\Phi }(𝐫)`$ and $`\sigma (𝐫)`$ are obtained by differentiation of equation 6. Before presenting the first few terms of this correspondence, we define two quantities $`𝐀_m`$ and $`𝐁_m`$, which are attributable to the $`m^{\mathrm{th}}`$ multipoles of the mass that is, respectively, exterior and interior to the Einstein ring radius: $`𝐀_m`$ $`=`$ $`{\displaystyle \frac{b^{m2}}{m}}{\displaystyle _b^{\mathrm{}}}𝑑rr^{1m}{\displaystyle \frac{𝝈_m(r)}{\sigma _\mathrm{c}}}`$ (17) $`𝐁_m`$ $`=`$ $`{\displaystyle \frac{b^{m2}}{m}}{\displaystyle _0^b}𝑑rr^{1+m}{\displaystyle \frac{𝝈_m(r)}{\sigma _\mathrm{c}}}.`$ (18) With these definitions, the correspondence between the higher multipoles may be written compactly. In general, for $`m1`$, $$𝐅_{mt}\frac{1}{b^2}b^t\frac{d^t}{dr^t}𝚽_m(r)|_{r=b}.$$ (19) Explicitly, the first four terms of the Taylor expansion are: $`𝐌_m^{\mathrm{sum}}𝐅_{m0}{\displaystyle \frac{1}{b^2}}𝚽_m(b)`$ $`=`$ $`𝐀_m+𝐁_m`$ (20) $`m𝐌_m^{\mathrm{diff}}𝐅_{m1}{\displaystyle \frac{1}{b^2}}b{\displaystyle \frac{d}{dr}}𝚽_m(r)|_{r=b}`$ $`=`$ $`m(𝐀_m𝐁_m)`$ (21) $`𝐅_{m2}{\displaystyle \frac{1}{b^2}}b^2{\displaystyle \frac{d^2}{dr^2}}𝚽_m(r)|_{r=b}`$ $`=`$ $`m(m1)𝐀_m+m(m+1)𝐁_m{\displaystyle \frac{2𝝈_m(b)}{\sigma _\mathrm{c}}}`$ (22) $`𝐅_{m3}{\displaystyle \frac{1}{b^2}}b^3{\displaystyle \frac{d^3}{dr^3}}𝚽_m(r)|_{r=b}`$ $`=`$ $`m(m1)(m2)𝐀_mm(m+1)(m+2)𝐁_m`$ (23) $`+{\displaystyle \frac{2𝝈_m(b)}{\sigma _\mathrm{c}}}2b{\displaystyle \frac{d}{dr}}{\displaystyle \frac{𝝈_m(r)}{\sigma _\mathrm{c}}}|_{r=b}.`$ Since the parameter of the $`t=0`$ term is the sum of the exterior and interior multipoles $`𝐀_m`$ and $`𝐁_m`$, we label it $`𝐌_m^{\mathrm{sum}}`$ rather than $`𝐅_{m0}`$. Likewise, the parameter of the $`t=1`$ term, labeled $`𝐌_m^{\mathrm{diff}}`$, is the difference between the exterior and interior multipoles. Higher-order terms ($`𝐅_{mt}`$ for $`t2`$) depend on $`𝐀_m`$ for $`tm`$, and on $`𝐁_m`$ for all $`t`$, as well as on the behavior of $`𝝈_m(r)`$ near $`r=b`$. In the next section it will be shown how to modify this expansion in order to separate explicitly the effects of $`𝐀_m`$ and $`𝐁_m`$ from the contributions that depend on $`𝝈_m(r)`$ in the vicinity of $`r=b`$. ### 3.3 Separation of internal and external contributions The physical meaning of our parameterization becomes clearer if $`𝐀_m`$ and $`𝐁_m`$ are used directly as parameters, instead of $`𝐌_m^{\mathrm{sum}}`$ and $`𝐌_m^{\mathrm{diff}}`$. Furthermore, the sum over $`t`$ of all the terms involving $`𝐀_m`$ and $`𝐁_m`$ can be calculated exactly, leaving only the effect of the mass near the ring radius left in the Taylor series. When this is done, the resulting “modified” multipole-Taylor expansion is: $`{\displaystyle \frac{1}{b^2}}\mathrm{\Phi }(r,\theta )`$ $`=`$ $`\mathrm{const}+\rho +{\displaystyle \frac{1}{2}}\rho ^2f_2+{\displaystyle \underset{t=3}{\overset{\mathrm{}}{}}}{\displaystyle \frac{1}{t!}}\rho ^tf_t`$ $`+{\displaystyle \underset{m=1}{\overset{\mathrm{}}{}}}\left\{\left({\displaystyle \frac{r}{b}}\right)^m𝐀_m\left({\displaystyle \frac{r}{b}}\right)^m𝐁_m+{\displaystyle \underset{t=2}{\overset{\mathrm{}}{}}}{\displaystyle \frac{1}{t!}}\rho ^t𝐆_{mt}\right\}(\widehat{𝐱}\mathrm{cos}m\theta +\widehat{𝐲}\mathrm{sin}m\theta )`$ In this expression, the terms $`𝐆_{mt}`$ are the portions of $`𝐅_{mt}`$ that are left when the contributions proportional to $`𝐀_m`$ and $`𝐁_m`$ are subtracted out. The generic form for $`𝐆_{mt}`$ is ($`m1`$, $`t2`$): $$𝐆_{mt}=\frac{1}{b^2}b^t\frac{d^t}{dr^t}𝚽_m^{\mathrm{near}\mathrm{ring}}(r)|_{r=b}$$ (25) where $$𝚽_m^{\mathrm{near}\mathrm{ring}}(r)=\frac{1}{m}\left(\frac{r}{b}\right)^m_r^b\left(\frac{r^{}}{b}\right)^{1m}\frac{dr^{}}{b}\frac{𝝈_m(r)}{\sigma _\mathrm{c}}\frac{1}{m}\left(\frac{r}{b}\right)^m_b^r\left(\frac{r^{}}{b}\right)^{1+m}\frac{dr^{}}{b}\frac{𝝈_m(r)}{\sigma _\mathrm{c}}$$ (26) Explicitly, the first four terms in the Taylor expansion are: $`𝐆_{m0}`$ $`=`$ $`0`$ (27) $`𝐆_{m1}`$ $`=`$ $`0`$ (28) $`𝐆_{m2}`$ $`=`$ $`2{\displaystyle \frac{𝝈_m(b)}{\sigma _\mathrm{c}}}`$ (29) $`𝐆_{m3}`$ $`=`$ $`2b{\displaystyle \frac{d}{dr}}{\displaystyle \frac{𝝈_m(r)}{\sigma _\mathrm{c}}}|_{r=b}2{\displaystyle \frac{𝝈_m(b)}{\sigma _\mathrm{c}}}`$ (30) We see that these parameters depend only on the distribution of the deflector mass near the ring radius. In particular, $`𝐆_{m2}`$ gives the $`m^{\mathrm{th}}`$ multipole moment of the matter located at the ring radius itself. ### 3.4 Degeneracies of the lens equation Certain parameters in the multipole-Taylor expansion cannot be constrained by observations of gravitational lensing, due to the so-called “degeneracies” of the lens equation. A degeneracy, as explained by Gorenstein, Falco & Shapiro (1988), is a family of distinct lens potentials and source distributions that all produce the same image configuration. Thus the observation of a given image configuration does not permit the actual lens potential and source distribution to be deduced, but rather only the degenerate family to which they belong. Breaking the degeneracy requires either additional assumptions or direct observations of the deflector itself (such as velocity dispersion measurements; see Falco, Gorenstein & Shapiro (1985)). The unconstrainable parameters must be fixed during the minimization of the error function $`\chi ^2`$, or else convergence is impossible. This section will identify those parameters and the fixed values we have chosen for them. The first degeneracy involves the choice of origin. From equations 18 and 9 follows, for $`m=1`$, $$𝐁_1=\frac{1}{b^3}_0^b𝑑rr^2\frac{𝝈_1(r)}{\sigma _\mathrm{c}}=\frac{𝐱_{\mathrm{com}}}{b},$$ (31) where $`𝐱_{\mathrm{com}}`$ is the center of mass of all the mass interior to $`b`$. This parameter changes with the choice of origin, but obviously the arbitrary choice of origin cannot affect the image configuration. Therefore one may allow either $`𝐁_1`$ or the location of the origin to vary, but not both simultaneously. We decided to impose $`𝐁_1=0`$ and allow the location of the origin to vary during the minimization procedure. This choice forces the center of the interior ($`r<b`$) mass to be at the origin, thereby simplifying the physical interpretation of the higher multipoles. The second degeneracy is the “prismatic” degeneracy: any term of the form $`𝐜𝐫`$ in the potential, where $`𝐜`$ is a constant vector, is unconstrainable (Falco, Gorenstein & Shapiro (1985); Gorenstein, Falco & Shapiro (1988)). The contribution of the exterior dipole $`𝐀_1`$ to the potential (equation 3.1 or equation 3.3) is the additive term $`b𝐀_1𝐫`$, which is precisely a prismatic-degenerate term. Consequently, we may arbitrarily set $`𝐀_1`$ to zero during the model-fitting procedure. (If the expansion in equation 3.1 is used, rather than the “modified” expansion with internal and external contributions separated out, then the center-of-mass degeneracy and the prismatic degeneracy can be taken into account by omitting the $`𝐌_1^{\mathrm{sum}}`$ and $`𝐌_1^{\mathrm{diff}}`$ terms, which is equivalent to setting both $`𝐀_1`$ and $`𝐁_1`$ to zero.) The last and most problematic degeneracy is the “mass-sheet degeneracy” (Falco, Gorenstein & Shapiro (1985); Gorenstein, Falco & Shapiro (1988)). Two potentials $`\mathrm{\Phi }(𝐫)`$ and $`\mathrm{\Phi }^{}(𝐫)`$ cannot be distinguished by the observation of lensed images if they are related by the transformation: $$\mathrm{\Phi }^{}(𝐫)=(1\kappa )\mathrm{\Phi }(𝐫)+\frac{\kappa r^2}{2}.$$ (32) This is equivalent to reducing the surface mass density by a factor $`(1\kappa )`$ and adding a sheet of uniform mass density $`\kappa \sigma _\mathrm{c}`$: $$\sigma ^{}(𝐫)=(1\kappa )\sigma (𝐫)+\kappa \sigma _\mathrm{c}.$$ (33) The corresponding relations between the multipole-Taylor parameters of the original and transformed potential are: $`f_0^{}=(1\kappa )f_0+{\displaystyle \frac{1}{2}}\kappa =\mathrm{const}^{}`$ (36) $`f_1^{}=f_1=1`$ $`f_2^{}=(1\kappa )f_2+\kappa `$ $`t3:`$ $`f_t^{}=(1\kappa )f_t`$ (41) $`𝐀_m^{}=(1\kappa )𝐀_m`$ $`𝐁_m^{}=(1\kappa )𝐁_m`$ $`𝐆_{mt}^{}=(1\kappa )𝐆_{mt}`$ The constant offset has been altered, but this is immaterial since adding a constant to the potential has no effect on light deflection or differential time delays. The transformation must leave the ring radius unchanged, so the coefficient of $`\rho `$ in the monopole part of the potential is unchanged. The quadratic monopole term, $`f_2`$, which depends on the surface mass density at the ring radius (Eq. 15) has been changed, because the surface mass density at the ring radius has been changed. All the coefficients of the other terms in the model have been scaled by a factor $`1\kappa `$. These simultaneous adjustments of all the $`f_t`$ ($`t2`$), $`𝐀_m`$, $`𝐁_m`$, and $`𝐆_{mt}`$ parameters leave the model predictions unchanged. Therefore one of these parameters should be fixed in some manner during the model-fitting procedure to permit convergence of the minimization algorithm. We note that the transformation of equation 32 does affect the time delays, $$\mathrm{\Delta }t^{}=(1\kappa )\mathrm{\Delta }t.$$ (42) This however does not permit observations of the time delays to be used to determine the scaling factor $`(1\kappa )`$, since a rescaling of the time delay will be confused with a rescaling of the Hubble parameter. We discuss this issue further in section 5.2. To break the degeneracy, and permit convergence of the minimization algorithm, we may arbitrarily select a value of $`f_2`$; by equation 15, this is equivalent to an arbitrary choice of $`\sigma _0(b)`$. We find it convenient to fix $`f_2^{}=0`$, which happens to be correct for an isothermal sphere, for which $`\sigma _0(b)=\sigma _\mathrm{c}/2`$. If this is not true of the actual mass distribution, then the fitted (primed) amplitudes of all the $`m0`$ multipoles and the $`t3`$ monopole parameters will have been scaled by the same factor $`1\kappa `$ relative to the true (unprimed) amplitudes, where $$1\kappa =\frac{1}{2\left(1\frac{\sigma _0(b)}{\sigma _\mathrm{c}}\right)}$$ (43) These scaling factors are tabulated in Table 5 for several different choices of the actual monopole potential. For example, suppose the choice $`f_2^{}=0`$ is made, but the true potential is actually a point mass rather than an isothermal sphere. Then, in reality, $`\mathrm{\Phi }_0(𝐫)=b^2\mathrm{ln}r`$, and $`\sigma _0(b)=0`$. All of the best-fit parameters ($`f_t^{}`$, $`𝐀_m^{}`$, etc.) and the model time delays ($`\mathrm{\Delta }t^{}`$) are smaller than those ($`f_t`$, $`𝐀_m`$, $`\mathrm{\Delta }t`$, etc.) of the actual mass distribution, having been scaled by the same factor $`(1\kappa )=1/2`$. Note that this rescaling only affects the magnitudes of the multipoles; it does not affect their angles. It also does not affect the center-of-mass parameters $`g_x`$ and $`g_y`$, or the ring radius $`b`$. Once we have set $`f_2^{}=0`$, the next term in the monopole expansion, $`f_3^{}`$, is the lowest-order monopole parameter available to give information on the radial distribution of mass. After rescaling, its amplitude is related to the true surface mass density by $$f_3^{}=1b\frac{d}{dr}\mathrm{ln}\left(1\frac{\sigma _0(r)}{\sigma _\mathrm{c}}\right)|_{r=b}$$ (44) in contrast to equation 16. It depends on the fall-off of the surface mass density near the Einstein ring radius, and is most sensitive to that falloff when $`\sigma _0(b)`$ is close to $`\sigma _\mathrm{c}`$. For surface mass densities that do not increase with radius, $`f_3^{}<1`$. The values of $`f_3^{}`$ for various monopole potentials are given in Table 5. The lowest value for $`f_3^{}`$ listed in the table is $`1`$, for a mass sheet, but $`f_3^{}`$ can be even more negative for potentials with a core radius, or if $`\sigma _0(r)`$ drops abruptly near the ring radius. The mass-sheet degeneracy also affects the interpretation of the amplitude of the quadratic multipole parameter $`𝐆_{m2}^{}`$, whose fitted value is related to the true surface mass density by $$𝐆_{m2}^{}=\frac{𝝈_m(b)}{\sigma _\mathrm{c}\sigma _0(b)},$$ (45) in contrast to equation 29. Non-negativity of the surface mass density limits the amplitude of $`𝐆_{m2}^{}`$: $$|𝐆_{m2}^{}|\frac{2\sigma _0(b)}{\sigma _\mathrm{c}\sigma _0(b)}.$$ (46) This limit ranges from zero, for $`\sigma _0(b)=0`$, to $`+\mathrm{}`$, for $`\sigma _0(b)=\sigma _\mathrm{c}`$. For a mass distribution with a singular isothermal sphere monopole, the limit is $`|𝐆_{m2}^{}|2`$., For mass distributions more centrally concentrated than a singular isothermal sphere, the limit on $`|𝐆_{m2}^{}|`$ would be lower. The limit is only attained if all the mass at the ring radius is clumped into point-like perturbers in any of the $`m`$ allowed directions. In summary, after expanding the lens potential in a multipole-Taylor series, then explicitly separating the contributions from the internal and external multipoles, and then fixing several parameters to zero (as described above) because of the degeneracies in the lens equation, the final parameterized form of the potential is: $`{\displaystyle \frac{1}{b^2}}\mathrm{\Phi }^{}(r,\theta )`$ $`=`$ $`\rho +{\displaystyle \underset{t=3}{\overset{\mathrm{}}{}}}{\displaystyle \frac{1}{t!}}\rho ^tf_t^{}`$ $`{\displaystyle \underset{m=2}{\overset{\mathrm{}}{}}}\left\{(1+\rho )^m𝐀_m^{}+(1+\rho )^m𝐁_m^{}\right\}(\widehat{𝐱}\mathrm{cos}m\theta +\widehat{𝐲}\mathrm{sin}m\theta )`$ $`+{\displaystyle \underset{m=1}{\overset{\mathrm{}}{}}}{\displaystyle \underset{t=2}{\overset{\mathrm{}}{}}}{\displaystyle \frac{1}{t!}}\rho ^t𝐆_{mt}^{}(\widehat{𝐱}\mathrm{cos}m\theta +\widehat{𝐲}\mathrm{sin}m\theta ).`$ Here and elsewhere, whenever the distinction is important, we have used primes to identify fitted model parameters, reserving the non-primed symbols for the parameters that describe the actual gravitational potential. These two sets of parameters differ because of the arbitrary degeneracy-breaking choices described in this section. Equation 3.4 is the parameterization we used to compute best-fit models for MG J0414+0534 (although, in a few cases, we tried using the parameter $`𝐌_{m}^{\mathrm{sum}}{}_{}{}^{}`$ from the original multipole-Taylor series, equation 3.1, in place of $`𝐀_m^{}`$ and $`𝐁_m^{}`$). A summary of the physical significance of each term in this expansion is given in Table 2. Since the analytic forms of the first and second derivatives of the potential are used during the model-fitting procedure, these derivatives are presented in Tables 3 and 4. ### 3.5 Comparison to other parameterizations Any parameterized form of the lens potential $`\mathrm{\Phi }(𝐫)`$ can be compared to ours by expanding it in a modified multipole-Taylor series. In this section we carry out this procedure for a few forms for the lens potential that are widely used. Often the lens potential is taken to be a combination of a monopole term and a quadrupole term, just as Kochanek (1991) did. The particular form of each of these is usually prescribed by either physical preconceptions or ease of computation. In Table 5 we compare five possible monopole potentials. These all give rise to the same linear term which sets the Einstein ring radius. The first significant term that distinguishes them is therefore $`f_3^{}`$, since $`f_2`$ is not constrainable due to the mass-sheet degeneracy. If the data are not sufficient to distinguish among the various possibilities for $`f_3^{}`$, then there is little point in using more complicated forms for the monopole term, such as de Vaucouleurs radial profiles (Ellithorpe (1995)). There are three different quadrupole terms that are commonly used to accompany the monopole term. The first is an external quadrupole, of the form: $$\mathrm{\Phi }_Q(r,\theta )=\frac{\gamma }{2}r^2\mathrm{cos}2(\theta \theta _\gamma )$$ (48) This term is sometimes referred to as an “external shear.” In our scheme, this is equivalent to choosing a non-zero value of $`𝐀_2`$, with magnitude $`|𝐀_2|=\gamma /2`$ and angle $`\psi _{A_2}=\pi /2+\theta _\gamma `$. (Our notation has $`\psi _{A_2}`$ and $`\psi _{A_2}+\pi `$ as the angles to the mass excess, whereas $`\theta _\gamma `$ and $`\theta _\gamma +\pi `$ are the angles to the mass deficit.) The second is an internal quadrupole term, parameterized as: $$\mathrm{\Phi }_Q(r,\theta )=\frac{\gamma }{2}\frac{b^4}{r^2}\mathrm{cos}2(\theta \theta _\gamma ),$$ (49) which is equivalent, in our scheme, to choosing a non-zero $`𝐁_2`$ with magnitude $`|𝐁_2|=\gamma /2`$ and angle $`\psi _{B_2}=\theta _\gamma `$. Finally, there is the case of a “mixed” quadrupole, $$\mathrm{\Phi }_Q(r,\theta )=\frac{b\gamma }{2}r\mathrm{cos}2(\theta \theta _\gamma ),$$ (50) which is obtained by truncating the potential of a singular isothermal elliptical potential at the quadrupole term. To order $`\rho `$, this is equivalent to the choice $`|𝐀_2|=3\gamma /8`$, $`|𝐁_2|=\gamma /8`$, and $`\psi _{A_2}=\psi _{B_2}=\theta _\gamma `$. It is worth noting here that $`𝐀_2+𝐁_2`$ causes tangential image displacements for images near the ring radius, whereas $`𝐀_2𝐁_2`$ causes radial image displacements. The balance between the internal and external portions of the quadrupole therefore determines the radial displacement that accompanies the tangential displacements caused by the quadrupole moment of the mass distribution. The choice of a “mixed” quadrupole is essentially a particular choice for this ratio. ### 3.6 Effects of external perturbing masses If there is a deflecting object along the line of sight to the source besides the primary lensing galaxy, this secondary deflector can be modeled in one of two ways. It could be treated separately from the primary deflector, with extra parameters for its location and mass distribution. Alternatively, the parameterization of equation 3.4 could be used alone, so that the influence of the secondary perturber would be reflected in the values of the multipole moments. Since the location (or even the presence) of a perturbing object is not known a priori, it is useful to compute the effect on the parameters of the multipole-Taylor expansion that would be caused by a perturbing object far removed from the primary deflector. A perturbing point mass located further from the origin than any of the lens images contributes to the external multipole moments to all orders $`m`$, with amplitudes $$|𝐀_m|=\frac{1}{m}\left(\frac{b}{R}\right)^m\left(\frac{b_P}{b}\right)^2,$$ (51) where $`b`$ is the ring radius of the principal deflector, $`b_P`$ is the ring radius of the perturber, and $`R`$ is the distance to the perturber. This can be derived by expanding its delta-function surface-mass distribution as a multipole-Taylor series. The $`m=1`$ term has no effect, due to the prismatic degeneracy, so the dominant term is the quadrupole moment, which is often called the external shear. For small $`b/R`$, it should be adequate to represent the effect of the perturber by the first few multipoles. Conversely, the ratios of the fitted amplitudes $`A_m`$ (if they can be attributed solely to a perturbing influence) permit the distance to the perturber, and its strength, to be deduced. How do radial and tangential extent in an external perturber affect the external multipole amplitudes? Consider a perturber with its center of mass located at a distance $`R`$ from the center of the principal deflector, with surface mass density uniform over a region of radial extent $`\mathrm{\Delta }R`$ and tangential extent $`R\mathrm{\Delta }\theta `$. Such a perturber contributes to the external multipole moments of the potential to all orders $`m`$ in the multipole expansion, with amplitudes $`|𝐀_m|={\displaystyle \frac{1}{m}}\left({\displaystyle \frac{b}{R}}\right)^m\left({\displaystyle \frac{b_P}{b}}\right)^2\{\text{}1`$ $`+`$ $`{\displaystyle \frac{m(m+1)}{24}}\left({\displaystyle \frac{\mathrm{\Delta }R}{R}}\right)^2{\displaystyle \frac{m(m+1)}{24}}\left(\mathrm{\Delta }\theta \right)^2`$ $`+`$ $`𝒪\left({\displaystyle \frac{\mathrm{\Delta }R}{R}}\right)^4+𝒪\left(\mathrm{\Delta }\theta \right)^4+𝒪\left(\mathrm{\Delta }\theta \right)^2𝒪\left({\displaystyle \frac{\mathrm{\Delta }R}{R}}\right)^2\text{}\},`$ The net effect, for perturbers with similar radial and tangential extents ($`\mathrm{\Delta }R=R\mathrm{\Delta }\theta `$) is that there is no effect through $`𝒪\left(\frac{\mathrm{\Delta }R}{R}\right)^3`$. Therefore, using the formulae for a point mass perturbation (eq. 51) should cause no problem for a perturber that is located far away as compared to its extent. Unfortunately, for a nearby extended perturber, the interpretation of the fitted external multipole amplitudes is less simple. ## 4 Application to MG J0414+0534 A model of a gravitational lens consists of two parts: a model of the surface brightness of the source, as it would appear in the absence of lensing, and a model of the lens potential. An ideal model reproduces the observed image of the system, pixel for pixel, by mapping the source surface brightness through the lens potential and then convolving with the detector response. This is the aim of modeling techniques such as LensClean (Kochanek & Narayan (1992); Ellithorpe, Kochanek & Hewitt (1996)), which use the information from every pixel to constrain the model. It is computationally faster to employ a much smaller number of constraints that capture the most important features of the observed image. Multiple images of nearly-pointlike sources, such as the 16 components seen in MG J0414+0534, are succinctly described by the locations and flux densities of the pointlike components. With this simplification, called “point modeling,” a much wider range of models can be explored. We chose to use only the positions of the 16 components of MG J0414+0534, and not their flux densities, as model constraints. There were two reasons for this. One, the relative uncertainties in the fluxes are much larger than the relative uncertainties in the centroid positions. Therefore the inclusion of flux information would not contribute much to the error statistic $`\chi ^2`$. Two, there are significant discrepancies between the radio measurements and optical measurements of the flux density ratios of components in MG J0414+0534. This discrepancy has been variously attributed to microlensing (Witt, Mao & Schechter (1995)), variable extinction (McLeod et al. (1998)), and/or substructure in the lensing galaxy (Mao & Schneider (1998)). Some of these effects may not be significant for radio observations, but we chose not to use flux density information at all. ### 4.1 Model-fitting algorithm Our constraints consist of the observed positions $`𝐫_{k\alpha }`$, where the index $`\alpha `$ runs over the 4 different images (A1, A2, B, and C) and the index $`k`$ runs over the 4 components (p, q, r, and s) of each image. The model consists of presumed source positions $`𝐬_k`$ for each component, along with the mapping $`𝐫_\alpha (𝐬_k)`$ provided by the lens equations (eq. 3) using our expansion of the potential (eq. 3.4) truncated to a prescribed order. Assuming that the observational errors are Gaussian, the maximum-likelihood values of the parameters can be determined by minimizing the familiar chi-squared statistic, $$\chi ^2=\underset{k=1}{\overset{4}{}}\underset{\alpha =1}{\overset{4}{}}(𝐫_\alpha (𝐬_k)𝐫_{k\alpha })𝐒_{k\alpha }^1(𝐫_\alpha (𝐬_k)𝐫_{k\alpha }),$$ (53) where $`𝐒_{k\alpha }`$ is the error covariance matrix. Because the lens mapping is much easier to apply in the direction $`𝐫𝐬`$ than the reverse direction, it is of great computational advantage to compute an approximate value of $`\chi ^2`$ by evaluating the model errors in the source plane rather than the image plane (Kayser et al. (1989)). By Taylor expansion, the displacement between the model source position and the actual source position is $$𝐬_k𝐬(𝐫_{k\alpha })=𝐌^1(𝐫_{k\alpha })(𝐫_\alpha (𝐬_k)𝐫_{k\alpha })+\left((𝐫_\alpha (𝐬_k)𝐫_{k\alpha })\mathbf{}_𝐫𝐌^1(𝐫_{k\alpha })\right)(𝐫_\alpha (𝐬_k)𝐫_{k\alpha })+\mathrm{},$$ (54) where $`𝐌^\mathrm{𝟏}(𝐫_{k\alpha })`$ is the inverse magnification matrix, $$M_{}^{1}{}_{ij}{}^{}(𝐫)=\delta _{ij}\frac{}{r_i}\frac{}{r_j}\mathrm{\Phi }(𝐫).$$ (55) For good models, the difference between the modeled and observed image-plane positions, $`𝐫_\alpha (𝐬_k)𝐫_{k\alpha }`$, is small, so the higher order terms in the expansion may be neglected. Equivalently, the change in magnification between the observed image location and the model-predicted image location is assumed to be negligible. The resulting “source-plane” approximation for $`\chi ^2`$ is: $$\chi ^2\chi _s^2=\underset{k=1}{\overset{4}{}}\underset{\alpha =1}{\overset{4}{}}(𝐬_k𝐬(𝐫_{k\alpha }))𝐌(𝐫_{k\alpha })𝐒_{k\alpha }^1𝐌(𝐫_{k\alpha })(𝐬_k𝐬(𝐫_{k\alpha }))$$ (56) This approximation is useful because it is valid near the chi-squared minimum, and we are unconcerned with the behavior of the function far from the minimum. The global minimum of both $`\chi ^2`$ and $`\chi _s^2`$ is zero, which occurs when the model exactly reproduces the observation. We expect that even with noise and measurement error, $`\chi _s^2`$ has a global minimum corresponding to that of the true chi-squared, $`\chi ^2`$, and that no lower minimum is introduced by this approximation. This approximation fails for very poor deflector models that do not reproduce the image locations at all — but this is of little concern, since the high $`\chi _s^2`$ would cause the model to be rejected anyways. The approximation can also fail if the error in the observed image locations is large enough to encompass a region in which the magnification matrix varies significantly. Because of this danger, after the minimum of $`\chi _s^2`$ was found for each model, we computed the true $`\chi ^2`$ to check the source-plane approximation. For models that adequately satisfied the observational constraints (as described below), $`\chi _s^2`$ typically differed from $`\chi ^2`$ by only $`0.20.6\%`$, and at most $`1.5\%`$. Since this is much less that the $`\chi ^2`$ increment used to find confidence limits on model parameters, the source-plane approximation introduced no appreciable error. Since the first and second derivatives of our parameterized potential are available in analytic form, the magnification matrix and thus $`\chi _s^2`$ are easy to compute. These derivatives are listed in Tables 3 and 4. Furthermore, since $`\chi _s^2`$ is quadratic in the source positions $`𝐬_k`$, the optimal source positions are easily computed for given values of the model parameters. Consequently, the numerical minimization need not include a search through the source positions in addition to the parameters of the deflector model. This reduction in the number of dimensions of the search space permits a vast computational speed-up. (If fluxes also are used as model constraints, a corresponding source-place approximation can still be made. The resulting $`\chi _s^2`$ is quadratic in both model source positions and model fluxes, so the optimal model positions and fluxes can still be found analytically \[Trotter (1998)\].) To perform the minimization of $`\chi _s^2`$ we employed a variant of simulated annealing described in Press et al. (1992). Simpler methods, such as a straightforward downhill walk via the Powell direction set method, or the downhill simplex method (Press et al. (1992)), encountered difficulties with local minima. The initial value for the center of mass was set to the galaxy position observed in the HST images of Falco, Lehár & Shapiro (1997). The initial value of the ring radius $`b`$ was taken from the previous best-fit models of Ellithorpe (1995), and the initial values of all $`m1`$ multipoles were set to zero. To explore the region of parameter space near these physically-motivated starting values, we used a moderately low starting temperature equal to 1% of the initial value of $`\chi _s^2`$. High starting temperatures allow the possibility of escaping the physically reasonable portion of parameter space and becoming trapped in deep and distant local minima. The higher multipole parameters were represented by their Cartesian components, e.g. $`\widehat{𝐱}𝐀_2^{}`$ and $`\widehat{𝐲}𝐀_2^{}`$, rather than by amplitude and angle. The potential depends linearly on these parameters, and thus the chi-squared depends roughly quadratically on them (far from the minimum, at least), allowing for a more robust minimization. However, confidence limits were computed for the amplitude and angle of each multipole parameter, rather than its Cartesian components, because the amplitude-angle representation is more useful for visualizing the mass distribution, and because amplitudes are affected by the mass-sheet degeneracy but angles are not. For the value of $`\chi ^2`$ to be used to calculate confidence intervals for the model parameters, the estimates of the observational errors (as represented in the covariance matrix $`𝐒_{k\alpha }`$) must be accurate. Unfortunately, it is difficult to make accurate error estimates of VLBI centroid positions, because of the complicated and nonlinear process of deconvolution by “cleaning” and self-calibration. In addition, if there are magnification gradients across the image, the image centroid may not be exactly the image of the source centroid, even though the point-modeling approach assumes so. The separation between the image of the source-centroid and the centroid of the image is approximately the angular size of the image multiplied by the fractional change in the magnification over the extent of the image. (The expression for the discrepancy is given by Trotter (1998), along with a correction to $`\chi _s^2`$ to account for it. We did not use this correction because it requires the calculation of the third derivatives of the potential and accurate estimates of the intrinsic source size.) For these reasons we report the results using three different methods to estimate the positional error in each component. The first, a crude upper limit, is the image size convolved with the VLBA beamwidth, which we call “fit-size errors.” The second estimate for the positional error is a lower limit: the statistical error in the centroid position due to thermal noise in the map. We report this as “statistical error.” The third estimate is the quadrature sum of the statistical error and the width of the deconvolved image (the intrinsic component size). This “stat-width error” makes some allowance for magnification gradients as well as deconvolution error. We believe the stat-width estimate to be the most accurate of the three estimates, but since this judgment is not rigorous we report the results for $`\chi ^2`$ using all three estimates. Confidence limits on each model parameter were determined using the “stat-width error,” by stepping the parameter away from the $`\chi _s^2`$ minimum, while minimizing over all other parameters, until the $`\mathrm{\Delta }\chi ^2`$ appropriate for 68.3% confidence limits was obtained. ### 4.2 Model results Table 6 summarizes the goodness-of-fit for a variety of model potentials. For each model the number of parameters and number of degrees of freedom are listed, along with the minimum $`\chi ^2`$ obtained using each of the three different estimates of positional errors. In this section we review these results. Before using the modified multipole-Taylor expansion, we tested three simpler parameterizations that have been used in previous attempts to model MG J0414+0534 based on lower-resolution radio and optical data. These simpler potentials were a singular isothermal sphere plus external shear (SIS+XS), a point mass with external shear (PM+XS), and a singular isothermal elliptical potential truncated at the quadrupole moment (SIEP) (see e.g. Hewitt et al. (1992), Falco, Lehár & Shapiro (1997), Ellithorpe (1995)): $`\mathrm{\Phi }_{\mathrm{SIS}+\mathrm{XS}}(r,\theta )`$ $`=`$ $`br+{\displaystyle \frac{1}{2}}\gamma r^2\mathrm{cos}2(\theta \theta _\gamma )`$ (57) $`\mathrm{\Phi }_{\mathrm{PM}+\mathrm{XS}}(r,\theta )`$ $`=`$ $`b^2\mathrm{ln}r+{\displaystyle \frac{1}{2}}\gamma r^2\mathrm{cos}2(\theta \theta _\gamma )`$ (58) $`\mathrm{\Phi }_{\mathrm{SIEP}}(r,\theta )`$ $`=`$ $`br{\displaystyle \frac{b\gamma }{2}}r\mathrm{cos}2(\theta \theta _\gamma )`$ (59) The results for these 5-parameter models are shown in the top three lines of Table 6. They are all very poor fits. Qualitatively they fail to reproduce the detailed VLBI structure of the four components. Quantitatively, even for the upper-limit (fit-size) errors, the $`\chi ^2`$ is more that $`3\times 10^3`$ standard deviations away from the value that would be expected for observations matching these models. The rest of the models in Table 6 are multipole-Taylor expansions truncated in various ways. They are labeled with symbols indicating the terms that are present in the expansion. The dominant terms in the multipole-Taylor expansion — the only terms that cause shifts in the image positions for images at the ring radius — are the first (linear in $`\rho `$) monopole term, $`b`$, and the first two (constant and linear in $`\rho `$, or external and internal) $`m2`$ multipole terms, $`𝐀_m`$ and $`𝐁_m`$. Accordingly, all of the models included the $`b`$ term, which sets the Einstein ring radius, as well as the internal quadrupole ($`𝐁_2`$) and external shear ($`𝐀_2`$) terms which account for ellipticity of the deflector mass distribution and the dominant effects of any external perturber. The model containing these three terms and no others is given the schematic name $`b+𝐀_2+𝐁_2`$, and is the fourth model listed in Table 6. It differs from the SIS+XS model only by the addition of the internal quadrupole $`𝐁_2`$ term. The addition of this second quadrupole term causes a vast improvement in the fit; the minimum $`\chi ^2`$ is lowered by two orders of magnitude, although the model is still not in formal agreement with the data. All of the other models are labeled with a “+” sign and the terms they contain in addition to the three terms $`b`$, $`𝐀_2`$, and $`𝐁_2`$. The next batch of models, as indicated in Table 6, each include only one term in addition to these three. The next most significant terms in each of the $`m=0`$ through $`m=4`$ multipoles were tried. The $`m=3`$ multipoles were considered because they are the next terms in which effects of an external perturber would appear, and would also account for any lopsidedness in the mass distribution of the lens galaxy. The $`m=4`$ multipoles were considered to account for diskiness or boxiness of the lens galaxy. Higher multipoles, $`m5`$, were not tried, as there was no physical reason to expect them to be significant. Also tried were the mixed-internal-and-external terms, $`𝐌_3^{\mathrm{sum}}`$ and $`𝐌_4^{\mathrm{sum}}`$, of the original multipole-Taylor expansion, equation 3.1. Using $`𝐌_m^{\mathrm{sum}}`$ instead of $`𝐀_m`$ or $`𝐁_m`$ strikes a different balance between the internal and external contributions to the $`m^{\mathrm{th}}`$ multipole. In all cases the fits were an improvement over the three-term model, but the best results were obtained by adding either an external octupole $`𝐀_3`$, or a mixed-internal-and-external octupole $`𝐌_3^{\mathrm{sum}}`$. These two models fit the image positions well enough to satisfy the fit-size (upper limit) position errors, though not well enough to satisfy the two tighter error estimates. For both of these models, the angle of the external quadrupole ($`74.4\mathrm{°}\pm 0.2\mathrm{°}`$ E of N for $`+𝐀_3^{}`$; $`74.1\mathrm{°}\pm 0.2\mathrm{°}`$ for $`+𝐌_{3}^{\mathrm{sum}}{}_{}{}^{}`$) is consistent with the optical isophote angle of the deflector as observed in the WFPC2 image of Falco, Lehár & Shapiro (1997) ($`71\mathrm{°}\pm 5\mathrm{°}`$). For the $`+𝐀_3`$ model, the direction of the internal quadrupole $`𝐁_2^{}`$ ($`75.4\mathrm{°}\pm 0.5\mathrm{°}`$ E of N) also agrees with the WFPC2 optical isophotes. The best-fit parameters of this 9-parameter model ($`+𝐀_3`$) are displayed pictorially in Figure 4. The $`+𝐌_3^{\mathrm{sum}}`$ model has a somewhat smaller value of $`\chi ^2`$, although in this case the internal quadrupole $`𝐁_2^{}`$ ($`87.4\mathrm{°}\pm 0.9\mathrm{°}`$ E of N) and the isophotes are misaligned by $`16\mathrm{°}`$. To each of these promising models, $`+𝐀_3`$ and $`+𝐌_3^{\mathrm{sum}}`$, was added the next radial term in each $`m4`$ multipole, one at a time, as indicated in the third block of Table 6. These models had 10 or 11 parameters, depending on whether the additional term was monopole or not. For the sake of comparison, a model with 11 parameters but employing $`m=4`$ multipoles instead of $`m=3`$ multipoles was also tried (and fared very poorly). The models including the external octupole $`𝐀_3`$ outperformed the models including the mixed octupole $`𝐌_3^{\mathrm{sum}}`$. The model with the lowest $`\chi ^2`$ was $`+𝐀_3+𝐀_4`$ (external $`m=3`$ and $`4`$ multipoles), followed by the model $`+𝐀_3+𝐁_3`$ (external and internal $`m=3`$ multipoles).<sup>1</sup><sup>1</sup>1 The model $`+𝐀_3+𝐆_{22}`$ produced the unphysical value $`|𝐆_{22}^{}|=0.5`$. This would require that at least $`25\%`$ of the mass in a narrow annulus at $`r=b`$ be involved in driving the quadrupole term — or more, if mass is not concentrated into points. (This assumes that the mass of the deflector is not more extended than a singular isothermal sphere.) See equation 46. To produce sizable image displacements, $`|𝐆_{22}^{}|`$ must be large because it must overcome the small factor $`\rho `$. This term is apparently compensating for the internal and external quadrupoles, which are displaced relative to their orientations in the $`+𝐀_3`$ model. Both of these models oversatisfy even the stat-width error estimates (indicating, perhaps, that these error estimates may be too large), though neither model is formally consistent with the lower-limit error estimates. The best-fit parameters for the most successful 11-parameter model, $`+𝐀_3+𝐀_4`$, are shown pictorially in Figure 5. The implications of the success of this and other models will be considered in the next section. The lowest-order term in the multipole-Taylor expansion that is both constrainable and sensitive to the radial distribution of mass is $`f_3^{}`$. In the next round of modeling we added the term $`f_3^{}`$ to each of the models of the previous group, in order to determine whether the radial dependence of the potential could be usefully constrained. The results are shown in the last block of Table 6. The fitted parameter values have large error ranges (e.g. $`f_3^{}=1.13_{0.52}^{+0.39}`$ for the model $`+𝐀_3+𝐀_4+f_3`$) compared to the range of interesting values, which extends from $`f_3^{}=0`$ (singular isothermal sphere) to $`f_3^{}=1`$ (point mass). More problematic is that the value of this parameter depends sensitively on the presence or absence of the other multipole components in the model, with values ranging from $`9.7`$ to $`2.3`$ for 12-parameter models that adequately satisfy the stat-width error estimates. It is clear that useful information on the radial profile of MG J0414+0534 is unavailable from this data. ## 5 Discussion ### 5.1 Implications for the mass distribution The best-fit model of the lens potential with 11 parameters included an $`m=3`$ external multipole and an $`m=4`$ external multipole. What are the implications of this success for the mass distribution of the deflector? To recapitulate, the explicit form of the model potential in this case is: $`\mathrm{\Phi }^{}(x,y)`$ $`=`$ $`b^2(\text{}\rho ((1+\rho )^2𝐀_2^{}+{\displaystyle \frac{1}{(1+\rho )^2}}𝐁_2^{})(\widehat{𝐱}\mathrm{cos}2\theta +\widehat{𝐲}\mathrm{sin}2\theta )`$ $`(1+\rho )^3𝐀_3^{}(\widehat{𝐱}\mathrm{cos}3\theta +\widehat{𝐲}\mathrm{sin}3\theta )(1+\rho )^4𝐀_4^{}(\widehat{𝐱}\mathrm{cos}4\theta +\widehat{𝐲}\mathrm{sin}4\theta )\text{})`$ where $`\theta `$ is measured north of west about the center of mass $`(g_x,g_y)`$ as given by $`\mathrm{tan}\theta =(yg_y)/(xg_x)`$, and the radial parameter $`\rho `$ is given by $$\rho =\left(\sqrt{(xg_x)^2+(yg_y)^2}b\right)/b.$$ (61) Table 7 contains a list of the best-fit parameters. In addition, Table 8 lists the image magnifications predicted by this best-fit model (which may be compared to the flux ratios in Table 1). Figure 3 shows both the observed and modeled image locations for each of the sixteen components of the VLBI map, along with the stat-width error ellipses. In this model, the center of mass of the mass distribution interior to $`r=b`$ is located at $`\mathrm{\Delta }\alpha =1\mathrm{}.0788\pm 0\mathrm{}.0020`$, $`\mathrm{\Delta }\delta =0\mathrm{}.6635\pm 0\mathrm{}.0012`$ relative to the correlation center at component A1p. Since the position of component Cp relative to the correlation center was $`\mathrm{\Delta }\alpha =1\mathrm{}.9454`$, $`\mathrm{\Delta }\delta =0\mathrm{}.3004`$ (with position errors negligible compared to those of the model galaxy, see Table 1) the model center of mass is $`\mathrm{\Delta }\alpha =0\mathrm{}.8666\pm 0\mathrm{}.0020`$, $`\mathrm{\Delta }\delta =0\mathrm{}.3631\pm 0\mathrm{}.0012`$ relative to Cp. This is in agreement with the optical centroid of the lens galaxy as observed by Falco, Lehár & Shapiro (1997) even though the optical position was not used as a modeling constraint. The observed lens galaxy position in $`R`$-band, relative to component C, was $`\mathrm{\Delta }\alpha =0\mathrm{}.90\pm 0\mathrm{}.05`$, $`\mathrm{\Delta }\delta =0\mathrm{}.32\pm 0\mathrm{}.05`$; the observed position in $`I`$-band was $`\mathrm{\Delta }\alpha =0\mathrm{}.86\pm 0\mathrm{}.05`$, $`\mathrm{\Delta }\delta =0\mathrm{}.36\pm 0\mathrm{}.05`$. As is apparent in Table 7 and Figure 5, the directions of mass excess implied by both the internal and external quadrupoles ($`𝐀_2^{}`$ and $`𝐁_2^{}`$) agree with the direction of the optical isophotes ($`71\mathrm{°}\pm 5\mathrm{°}`$) observed by Falco, Leár & Shapiro (1997). Interestingly, one of the directions of mass deficit implied by the external $`m=4`$ multipole, $`𝐀_4^{}`$, is also in agreement with the optical isophote angle. The eastern direction of the external octupole, $`𝐀_3^{}`$, is aligned with the external quadrupole moment within $`6\mathrm{°}`$, and is within $`10\mathrm{°}`$ of the observed isophote angle, although in neither case do the formal confidence regions overlap. It is possible that the alignments between the multipole angles and the optical isophotes are coincidences. If the isophote angle were selected at random, the chance of agreement with the direction of mass-excess indicated by $`𝐀_2^{}`$ would be $`6\%`$, given the quoted confidence ranges. The chance of agreement with the direction of either the mass excess or mass deficit implied by $`𝐀_4^{}`$ would be 27%. The chance that $`𝐀_3^{}`$ and the isophote angles would be as closely aligned as they are is 34%. Bearing this in mind, we entertain three speculations regarding the origin of the external multipoles in the best-fit model $`b+𝐀_2+𝐁_2+𝐀_3+𝐀_4`$: 1. All of the external multipoles are attributable to the mass distribution of the lens galaxy. This explains the alignments of the various multipole angles with the optical isophotes, but it implies that no external perturber (i.e. neither object X nor the group of galaxies to the southwest) contributes significantly to the features of the potential we have modeled. A singular isothermal elliptical potential would have a ratio of external to internal quadrupole of $`|𝐀_2|/|𝐁_2|=3`$, which is consistent with the value $`2.90\pm 0.17`$ obtained for this model. The ellipticity ($`1b/a`$) of the isopotential contours near the ring radius for the fitted model is $`ϵ_\mathrm{\Phi }=2|𝐀_2^{}+𝐁_2^{}|=0.120\pm 0.002`$. In contrast the ellipticity of a singular isothermal elliptical potential (SIEP) having an isodensity ellipticity of $`ϵ_G=0.20\pm 0.02`$ (equal to the mean ellipticity of the fitted isophotes of Falco, Lehár & Shapiro (1997)) would be $`ϵ_\mathrm{\Phi }=0.07`$. However, the non-zero $`𝐀_3^{}`$ implies that the outer galactic halo is asymmetric, with more mass concentrated near one end of the isophote axis than the other. Furthermore, the magnitude of $`𝐀_4^{}`$ ($`4.1\times 10^3`$) is an order of magnitude larger than the value it would assume for an singular isothermal elliptical potential ($`6\times 10^4`$). The direction of $`𝐀_4^{}`$ implies that the mass distribution is box-like, rather than disk-like as for a SIEP. 2. $`𝐀_2^{}`$ and $`𝐀_3^{}`$ indicate an external perturbing mass to the east. In this case the alignments of all multipoles with the isophotes are accidental. According to equation 51, a point mass with an Einstein radius of $`0.95\pm 0.05`$ arcseconds, located $`3.2\pm 0.2`$ arcseconds away, would supply approximately the appropriate values of $`𝐀_2^{}`$ and $`𝐀_3^{}`$, but would not account for $`𝐀_4^{}`$. No perturber of any kind is seen to the east in the optical images of Falco, Lehár & Shapiro (1997). 3. $`𝐀_3^{}`$ and $`𝐀_4^{}`$ indicate an external perturbing mass at about $`65\mathrm{°}\pm 4\mathrm{°}`$ N of W. In this case, as above, the alignments of all multipoles with the isophotes are accidental. A point mass $`2.24\pm 0.24`$ arcseconds away, with Einstein radius $`0.56\pm 0.10`$ arcseconds, would supply the proper $`|𝐀_3|`$ and $`|𝐀_4|`$, but would also make a significant contribution to $`𝐀_2`$. The residual component of $`𝐀_2^{}`$, which would presumably be due to the lensing galaxy, has magnitude 0.06 and direction $`1.5\mathrm{°}`$ or $`181.5\mathrm{°}`$ north of west. This is somewhat problematic because the residual quadrupole does not agree in angle with the optical isophotes, and its magnitude is uncomfortably large ($`|𝐀_2^{}|/|𝐁_2^{}|=3.8`$). However, there is an object seen in the optical image of Falco, Lehár & Shapiro (1997) about $`5\mathrm{}`$ away in the direction $`54\mathrm{°}`$ N of W. None of these speculations is entirely satisfactory. The first speculation attributes a peculiar shape to the galactic halo; the second invokes an external perturber that does not seem to be present; the third compels the galaxy to produce an unusual quadrupole moment. We favor the first interpretation, because it is hard to arrange for an external perturbation to produce $`𝐀_3^{}`$ without ruining the suggestive alignment of $`𝐀_2^{}`$ with the isophotes, but admit that this interpretation is debatable. This ambiguity of interpretation illustrates both the appeal and the frustration of the multipole-Taylor technique for modeling lens potentials. The technique makes few preconceptions about the shape of the potential, which in principle may lead to unanticipated discoveries about the mass distribution of the deflector, including the shape of the halo (of which little is presently known). However, precisely because of this mathematical generality, there is no determinative way to correlate features of the potential with observed astrophysical objects (the lens galaxy or external perturbers). Finally, we comment on the unconstrainability of the radial distribution of the mass monopole. It is not terribly surprising that we were unable to usefully constrain the parameter $`f_3^{}`$, the lowest-order parameter containing information about the radial distribution. A large change in $`f_3^{}`$ is needed to cause a small change in the radial positions of images near the ring radius. By contrast, even small values of $`𝐀_m^{}`$ or $`𝐁_m^{}`$ (for $`m2`$) cause shifts in the radial positions of images near the ring radius. In particular, varying $`𝐀_m^{}`$ and $`𝐁_m^{}`$ while leaving their sum unchanged affects the radial image displacements, but has little effect on the tangential image displacements (see Table 3). The radial image shifts caused by these multipole terms depend on angular position, whereas those caused by $`f_3^{}`$ do not. However, for lenses such as MG J0414+0534, in which there are only images at a few angular locations, the effects of $`f_3^{}`$ and of $`𝐀_m^{}`$ or $`𝐁_m^{}`$ may compete, with large changes in $`f_3^{}`$ compensating for small changes in $`𝐀_m^{}`$ or $`𝐁_m^{}`$. For systems with arcs or rings of lensed emission, information is available from a broader range of angles. The optical arc visible in MG J0414+0534 (Falco, Lehár & Shapiro (1997)) may further constrain MG J0414+0534’s angular multipole moments, especially if its location can be measured with the same precision as the VLBA measurements used in this paper. That the radial profile parameter, $`f_3^{}`$, is so difficult to determine is unfortunate. It would provide information on how the angularly-averaged surface mass density decreases with radius near the Einstein ring radius, which could be used help choose a model value for $`\sigma _0(b)`$. The quantity $`\sigma _0(b)`$ is not directly constrainable from lensing, but it does affects the predicted time delays between images, a topic discussed in the next section. ### 5.2 Implications for the time delays Each of the multiple images formed by a gravitational lens represents the source object at a different moment in its history. This is because the propagation time from source to observer is different for each image, due to the different path lengths and Shapiro delays experienced by the ray bundles composing each image. In particular, the time delay by which an image at $`𝐫_\alpha `$ lags that at $`𝐫_\beta `$ is, as computed by Narayan & Bartelmann (1996), $$\mathrm{\Delta }t_{\beta \alpha }=t(𝐫_\alpha )t(𝐫_\beta )=(1+z_l)\frac{D_LD_S}{D_{LS}}\left(\frac{1}{2}|\mathbf{}_{𝐫_\alpha }\mathrm{\Phi }(𝐫_\alpha )|^2\frac{1}{2}|\mathbf{}_{𝐫_\beta }\mathrm{\Phi }(𝐫_\beta )|^2\mathrm{\Phi }(𝐫_\alpha )+\mathrm{\Phi }(𝐫_\beta )\right)$$ (62) It is convenient to define a “dimensionless time delay” $`\mathrm{\Delta }\tau _{\beta \alpha }`$ which depends only on the modeled lens potential and requires no values for redshifts or cosmological parameters, $$\mathrm{\Delta }\tau _{\beta \alpha }=\frac{1}{2}|\mathbf{}_{𝐫_\alpha }\mathrm{\Phi }(𝐫_\alpha )|^2\frac{1}{2}|\mathbf{}_{𝐫_\beta }\mathrm{\Phi }(𝐫_\beta )|^2\mathrm{\Phi }(𝐫_\alpha )+\mathrm{\Phi }(𝐫_\beta )$$ (63) and which is related to the time delay by a conversion factor depending on the redshifts and cosmology, $$\mathrm{\Delta }t_{\beta \alpha }=(1+z_l)\frac{D_LD_S}{D_{LS}}\mathrm{\Delta }\tau _{\beta \alpha }.$$ (64) A measurement of the time delay between the flux variations in corresponding images and the lens redshift $`z_l`$, when combined with a model that predicts the dimensionless time delay, thereby amounts to a measurement of the combination of angular diameter distances $`D_LD_S/D_{LS}`$. Since the relation between angular diameter distance and redshift depends on the values of $`H_0`$, $`\mathrm{\Omega }_m`$, and $`\mathrm{\Omega }_\mathrm{\Lambda }`$, these cosmological parameters can be thereby constrained. The appeal of this well-known cosmological probe is that it does not rely on the usual intermediate distance indicators (Refsdal 1964, (1966)). One problem with this idea arises from the mass-sheet degeneracy, which was discussed in section 3.4. When the potential is transformed by the mass sheet degeneracy, the time delays between components are likewise transformed. Using the model potential of equation 3.4, the model dimensionless time delay, $`\mathrm{\Delta }\tau ^{}`$, is related to the true dimensionless time delay, $`\mathrm{\Delta }\tau `$, by $$\mathrm{\Delta }\tau _{\alpha \beta }^{}=\frac{1}{2\left(1\frac{\sigma _0(b)}{\sigma _\mathrm{c}}\right)}\mathrm{\Delta }\tau _{\alpha \beta }$$ (65) where $`\sigma _0(b)`$ is the angularly-averaged surface mass density at the Einstein ring radius. Unless the mass-sheet degeneracy can be resolved by determining $`\sigma _0(b)/\sigma _\mathrm{c}`$ in some independent fashion the time delays cannot be predicted, although the ratios of time delays between different image pairs can still be predicted. Although MG J0414+0534 has been extensively monitored at radio wavelengths, no flux variations have been observed that are large enough to permit an accurate time delay measurement (Moore & Hewitt (1997)). Nevertheless, this does not preclude radio or optical detections of time delays in the future, so it is important to understand how our models of MG J0414+0534 constrain the time delays. As discussed in the previous section, the best-fit model was $`b+𝐀_2+𝐁_2+𝐀_3+𝐀_4`$. We used this model to make our single best prediction for the time delays of MG J0414+0534, by computing the dimensionless time delays between the 4 images of the brightest component (p). The uncertainty in this particular model’s predicted time delay due to the uncertainty in the measured image positions was estimated in the following manner. We re-computed the time delay with each parameter (one at a time) adjusted to its maximum and minimum values allowed by the stat-width confidence limits, while minimizing the $`\chi ^2`$ over all the other parameters. The range between the highest and lowest time delays that were achieved during these parameter-by-parameter adjustments is our estimate for the uncertainty in the time delay for that particular model. However, we must also take into account the larger uncertainty in the predicted time delay caused by the uncertain choice of model. To estimate this uncertainty, we computed dimensionless time delays for a large subset of the models that were discussed in the previous section. For the 11-parameter and 12-parameter models, we included models which adequately fit the observations when using the stat-width error estimates ($`N_\sigma <3`$). The $`+𝐀_3+𝐌_4^{\mathrm{sum}}+f_3`$ and $`+𝐀_3+𝐁_4+f_3`$ models were excluded because, for these models, the entire confidence range for $`f_3^{}`$ lies above $`f_3^{}>2`$, well within the unphysical region $`f_3^{}>1`$. Likewise, the $`+𝐀_3+𝐆_{22}`$ model was not included because of its unphysically large value of $`|𝐆_{22}^{}|`$, as discussed in section 4.2. The results are shown in Table 9. Since these models make somewhat different predictions, any attempt to make a single prediction for the time delays must incorporate the uncertainty associated with the selection of a single model. Thus we have enlarged the error spread in our best predictions for the time delays to include the whole range of time delays predicted by all these models. The resulting predictions are: $`\mathrm{\Delta }\tau _{BA}^{}={\displaystyle \frac{1}{2\left(1\frac{\sigma _0(b)}{\sigma _\mathrm{c}}\right)}}\mathrm{\Delta }\tau _{BA}`$ $`=`$ $`1.828\times 10^{12}\left[1{}_{}{}^{+}\underset{\begin{array}{c}\mathrm{formal}\hfill \\ \mathrm{errors}\text{}\hfill \end{array}}{\underset{}{{}_{0.018}{}^{}{}_{}{}^{0.020}}}\pm \underset{\begin{array}{c}\mathrm{A1}\mathrm{A2}\hfill \\ \mathrm{difference}\text{}\hfill \end{array}}{\underset{}{0.014}}{}_{}{}^{+}\underset{\begin{array}{c}\mathrm{which}\hfill \\ \mathrm{model}\text{}\hfill \end{array}}{\underset{}{{}_{0.038}{}^{}{}_{}{}^{0.616}}}\right],`$ (72) $`\mathrm{\Delta }\tau _{AC}^{}={\displaystyle \frac{1}{2\left(1\frac{\sigma _0(b)}{\sigma _\mathrm{c}}\right)}}\mathrm{\Delta }\tau _{AC}`$ $`=`$ $`1.042\times 10^{11}\left[1\pm \underset{\begin{array}{c}\mathrm{formal}\hfill \\ \mathrm{errors}\text{}\hfill \end{array}}{\underset{}{0.017}}\pm \underset{\begin{array}{c}\mathrm{A1}\mathrm{A2}\hfill \\ \mathrm{difference}\text{}\hfill \end{array}}{\underset{}{0.002}}{}_{}{}^{+}\underset{\begin{array}{c}\mathrm{which}\hfill \\ \mathrm{model}\text{}\hfill \end{array}}{\underset{}{{}_{0.198}{}^{}{}_{}{}^{0.085}}}\right],`$ (79) $`\mathrm{\Delta }\tau _{BC}^{}={\displaystyle \frac{1}{2\left(1\frac{\sigma _0(b)}{\sigma _\mathrm{c}}\right)}}\mathrm{\Delta }\tau _{BC}`$ $`=`$ $`1.225\times 10^{11}\left[1\pm \underset{\begin{array}{c}\mathrm{formal}\hfill \\ \mathrm{errors}\text{}\hfill \end{array}}{\underset{}{0.017}}{}_{}{}^{+}\underset{\begin{array}{c}\mathrm{which}\hfill \\ \mathrm{model}\text{}\hfill \end{array}}{\underset{}{{}_{0.148}{}^{}{}_{}{}^{0.138}}}\right].`$ (84) Here the “formal errors” represent the uncertainty in the parameters of the best-fit model, the “A1-A2 difference” is half the time delay between images A1 and A2 (since the joint A1-A2 light curve would probably be used to measure time delays), and “which model” refers to the uncertainty due to choice of model. The factor $`2\left(1\sigma _0(b)/\sigma _\mathrm{c}\right)`$ represents the uncertainty in the time delay due to the mass-sheet degeneracy, which can only be relieved by obtaining a reliable value of $`\sigma _0(b)/\sigma _\mathrm{c}`$ from other observational or theoretical sources. For a potential with the radial profile of an isothermal sphere, $`2\left(1\sigma _0(b)/\sigma _\mathrm{c}\right)=1`$. To express the time delay as an actual number of days, the conversion factor of equation 64 must be used. For MG J0414+0534, which has $`z_l=0.9584`$ and $`z_s=2.639`$, this conversion factor takes the value $`\mathrm{\Delta }t/\mathrm{\Delta }\tau =6.794\times 10^{12}h_{75}^1`$ days, assuming the universe has $`\mathrm{\Omega }_m=1`$, $`\mathrm{\Omega }_\mathrm{\Lambda }=0`$, and $`H_0=75h_{75}`$ km/s/Mpc. The values of this conversion factor for some other choices of cosmological parameters are tabulated in Table 10. In all cases, the “filled-beam” approximation was used to compute the conversion factor, in which the universe is assumed to have a perfectly smooth distribution of matter. The presence of clumpiness would require the angular-diameter distances to be re-computed (see e.g. Fukugita et al. (1992)). A promising way to reduce the “which model” uncertainty is to measure the time delay ratio $`\mathrm{\Delta }t_{AB}/\mathrm{\Delta }t_{AC}`$. Predictions for this ratio are presented in Table 11 for various models. If this quantity is in accordance with the prediction of our best-fit model $`b+𝐀_2+𝐁_2+𝐀_3+𝐀_4`$, and it can be measured to within 3%, then it would exclude all the other models listed and the “which model” error would fall away. Even if the ratio could only be measured to within 18%, it would exclude all but one other model, which would shrink the “which model” uncertainty in $`\mathrm{\Delta }t_{AB}`$ to only 3.7% (from 62%). In this scenario, and using the conversion factor (eq. 64) for $`\mathrm{\Omega }_m=1`$, $`\mathrm{\Omega }_\mathrm{\Lambda }=0`$, we would predict H0=75km/s/Mpc12.42daysΔtBA[10.0180.020formalerrors+±0.014A1A2difference0.0370.000whichmodel+][2(1σ0(b)σc)]H0=75km/s/Mpc70.79daysΔtAC[1±0.017formalerrors±0.002A1A2difference0.1110.000whichmodel+][2(1σ0(b)σc)]H0=75km/s/Mpc83.23daysΔtBC[1±0.017formalerrors0.1000.000whichmodel+][2(1σ0(b)σc)]} Assuming the time delay ratio is measured to within 18% and agrees with our best-fit model. casessubscript𝐻075kmsMpc12.42daysΔsubscript𝑡𝐵𝐴delimited-[]plus-or-minus1superscriptsubscriptsubscriptsubscriptsuperscriptabsent0.0200.018formalerrorssubscript0.014A1A2differencesuperscriptsubscriptsubscriptsubscriptsuperscriptabsent0.0000.037whichmodeldelimited-[]21subscript𝜎0𝑏subscript𝜎csubscript𝐻075kmsMpc70.79daysΔsubscript𝑡𝐴𝐶delimited-[]plus-or-minus1subscript0.017formalerrorssubscript0.002A1A2differencesuperscriptsubscriptsubscriptsubscriptsuperscriptabsent0.0000.111whichmodeldelimited-[]21subscript𝜎0𝑏subscript𝜎csubscript𝐻075kmsMpc83.23daysΔsubscript𝑡𝐵𝐶delimited-[]plus-or-minus1subscript0.017formalerrorssuperscriptsubscriptsubscriptsubscriptsuperscriptabsent0.0000.100whichmodeldelimited-[]21subscript𝜎0𝑏subscript𝜎c Assuming the time delay ratio is measured to within 18% and agrees with our best-fit model. \left.\begin{array}[]{ccl}H_{0}&=&75\,{\rm km/s/Mpc}\,\frac{12.42\,\rm days}{\Delta t_{BA}}\,\left[1{}^{+}_{-}\underbrace{{}^{0.020}_{0.018}}_{\begin{array}[t]{@{}l@{}}{\scriptscriptstyle\rm formal}\\ {\scriptscriptstyle\rm errors\rule{0.0pt}{2.15277pt}}\end{array}}\pm\underbrace{0.014}_{\begin{array}[t]{@{}l@{}}{\scriptscriptstyle\rm A1-A2}\\ {\scriptscriptstyle\rm difference\rule{0.0pt}{2.15277pt}}\end{array}}{}^{+}_{-}\underbrace{{}^{0.000}_{0.037}}_{\begin{array}[t]{@{}l@{}}{\scriptscriptstyle\rm which}\\ {\scriptscriptstyle\rm model\rule{0.0pt}{2.15277pt}}\end{array}}\right]\left[{2\left(1-\frac{\sigma_{0}(b)}{{\sigma_{\rm c}}}\right)}\right]\\ H_{0}&=&75\,{\rm km/s/Mpc}\,\frac{70.79\,\rm days}{\Delta t_{AC}}\,\left[1\pm\underbrace{0.017}_{\begin{array}[t]{@{}l@{}}{\scriptscriptstyle\rm formal}\\ {\scriptscriptstyle\rm errors\rule{0.0pt}{2.15277pt}}\end{array}}\pm\underbrace{0.002}_{\begin{array}[t]{@{}l@{}}{\scriptscriptstyle\rm A1-A2}\\ {\scriptscriptstyle\rm difference\rule{0.0pt}{2.15277pt}}\end{array}}{}^{+}_{-}\underbrace{{}^{0.000}_{0.111}}_{\begin{array}[t]{@{}l@{}}{\scriptscriptstyle\rm which}\\ {\scriptscriptstyle\rm model\rule{0.0pt}{2.15277pt}}\end{array}}\right]\left[{2\left(1-\frac{\sigma_{0}(b)}{{\sigma_{\rm c}}}\right)}\right]\\ H_{0}&=&75\,{\rm km/s/Mpc}\,\frac{83.23\,\rm days}{\Delta t_{BC}}\,\left[1\pm\underbrace{0.017}_{\begin{array}[t]{@{}l@{}}{\scriptscriptstyle\rm formal}\\ {\scriptscriptstyle\rm errors\rule{0.0pt}{2.15277pt}}\end{array}}{}^{+}_{-}\underbrace{{}^{0.000}_{0.100}}_{\begin{array}[t]{@{}l@{}}{\scriptscriptstyle\rm which}\\ {\scriptscriptstyle\rm model\rule{0.0pt}{2.15277pt}}\end{array}}\right]\left[{2\left(1-\frac{\sigma_{0}(b)}{{\sigma_{\rm c}}}\right)}\right]\end{array}\right\}\rule{5.0pt}{0.0pt}\parbox{78.04842pt}{ Assuming the time delay ratio is measured to within $\sim$18\% and agrees with our best-fit model.} (85) ## 6 Conclusions Upon first seeing the rich sub-structure in each VLBI image of MG J0414+0534, we were hopeful that such a large body of precise constraints on the modeling potential would lead to tight constraints on the mass distribution of the deflector and the predicted time delays between images. We hoped that the mathematical generality of the multipole-Taylor expansion (with slight modifications) would allow us to draw such conclusions without contamination from (perhaps faulty) astrophysical preconceptions. These hopes were only partly fulfilled. Once the best-fit parameters in the expansion are determined, it is difficult to know from which astrophysical source they arise. For example, our best-fit model $`b+𝐀_2+𝐁_2+𝐀_3+𝐀_4`$ seems to imply that either the mass distribution in the lensing galaxy is asymmetric and somewhat quadrangular (boxy), or else that an unseen external perturber is partly responsible for the light deflection (as discussed in section 5.1). It is possible, however, that the values of the best-fit model parameters may guide future interpretations of observations for this system, by indicating the possible directions of perturbing masses. Predictions of time delays, while quite well-constrained for any particular model potential, are limited by the uncertainty in selecting one of several viable models. In other words, the predictions are not limited by the precision of the positional constraints, but rather by the ability to satisfy those constraints with several alternative truncations of the multipole-Taylor series. In the case of MG J0414+0534, we found that one way to relieve this crucial source of systematic error is to measure ratios of time delays, which are predicted to have different values by different models. Despite these limitations, which afflict all lens modeling techniques to date, the multipole-Taylor expansion does indeed seem to be an appropriate form for multiple-image gravitational lenses such as MG J0414+0534. The simplest three-term truncation reproduces the observed image configuration far better than previously-used simpler models, and successively higher terms improve the fit by incrementally smaller amounts. The radial profile parameter $`f_3^{}`$ could not be constrained, but this is likely to be the case for any point-modeling scheme applied to quadruple-image lenses. We believe the multipole-Taylor expansion could be usefully applied to other lenses in which the constraints occur at locations close to the Einstein radius, and the angular variation of the potential is expected to be fairly smooth. One serious problem with the efforts to date to determine $`H_0`$ and other cosmological parameters by measuring time delays is that each known gravitational lens has usually been modeled in an individual and idiosyncratic manner. The multipole-Taylor expansion is one candidate for a very general modeling technique that could be applied to all the time-delay lenses, so that the results of these efforts could be sensibly combined. We thank Paul Schechter for many valuable discussions concerning lens modeling in general and this analysis of MG J0414+0534 in particular. J. N. Winn thanks the Fannie and John Hertz Foundation for financial support. This work was supported by grant AST96-17028 from the National Science Foundation.
warning/0001/astro-ph0001144.html
ar5iv
text
# The Chemical Composition of Carbon Stars II: The J-type stars ## 1 Introduction The carbon content in the envelope of asymptotic giant branch (AGB) stars is believed to increase along the spectral sequence M$``$S$``$C during this phase of stellar evolution. The origin of this carbon enhancement is the mixing of He-burning products with matter from the convective envelope through the third dredge-up (TDU) mechanism which can happen after each thermal instability (pulse) of the He-shell (Iben & Renzini, 1983). The recurrence of TDU episodes leads to the creation of a carbon (C) star, defined as an AGB star with a C/O ratio higher than unity in the envelope. Among the C-stars there exists a significant group of stars ($`15\%`$) named J-type stars (Bouigue, 1954) showing strong <sup>13</sup>C-bearing molecule absorptions, which usually implies low <sup>12</sup>C/<sup>13</sup>C ratios ($`<15`$) (Lambert et al., 1986; Abia & Isern, 1997; Ohnaka & Tsuji, 1999). The location of J-stars in the above AGB spectral sequence is far from clear. In fact, some authors have located these stars in a different evolutive sequence from that of the ordinary carbon stars (e.g. Chen & Kwok 1993; Lorenz-Martins 1996), or even outside the AGB phase. J-stars have also been considered as the descendants of the late-R carbon stars, which have similar spectroscopic characteristics (Lloyd-Evans, 1986). Theoretically, it is not easy to obtain an AGB star with the chemical peculiarities presented by J-stars. Low <sup>12</sup>C/<sup>13</sup>C ratios can be obtained in current AGB star models of M$`4`$ M if hot H-burning takes place at the bottom of the convective envelope (the so-called hot bottom burning, HBB)(Lattanzio, 1999; Sackmann & Boothroyd, 1992). However, the performance of the CN-cycle at the same time destroys <sup>12</sup>C and, in consequence, the C/O ratio in the envelope is reduced and the star again becomes O-rich. Thus, a fine-tuning of the parameters of the AGB models (mass, mixing-length, mass-loss rate, metallicity, etc.), that determine the chemistry of the envelope, seems to be required to obtain a J-star. Mixing at the He-core flash has also been proposed as an alternative scenario to form J-stars. In this event an injection of carbon-rich material from the core into the hydrogen-rich shell may occur. The introduction of core material (<sup>12</sup>C and <sup>4</sup>He) into a proton-rich region yields enhanced <sup>12</sup>C and <sup>13</sup>C, with perhaps a small enhancement of <sup>14</sup>N (Deupree & Cole, 1983). The presence of strong $`\lambda 6708`$ Å Li I lines is frequent in J-stars. In a Li survey of galactic C-stars, Boffin et al. (1993) found that among 30 Li-rich stars in a sample of 250 C-stars $`50\%`$ are of J-type. This figure increases up to $`70\%`$ if the Li-rich phenomenon is considered only among the J-type C-stars in the survey. Although a good statistic has not yet been obtained, Li-rich J-type stars have also been found in the Magellanic Clouds (Brewer et al., 1996). Interestingly, envelope burning models can simultaneously produce Li-rich and <sup>13</sup>C-rich AGB stars in models with initial mass M$`4`$ M. However, observations indicate that the majority of C-stars in the galaxy are low-mass objects, M$`23`$ M (see, for example, Claussen et al. 1987). For this mass range no envelope burning has been found in any AGB model. On the other hand, another important consequence of the TDU episodes in the AGB phase is the enrichment of the envelope with s-process elements. These elements are believed to be synthesized during the period between thermal pulses, when a <sup>13</sup>C-pocket (formed in the intershell region) is burned radiatively and supplies the neutrons necessary to activate the s-process, via the <sup>13</sup>C($`\alpha `$,n)<sup>16</sup>O reaction (Straniero et al., 1995). At the next TDU, the synthesized s-nuclei are mixed into the envelope. Thus, if J-type stars owe their <sup>12</sup>C enhancement to the operation of the TDU, they should also show some s-process element enrichment. There are very few abundance analyses of J-type stars. The most extensive study is still that by Utsumi (1970). This pioneering study based, however, on low dispersion (5-14 Å/mm) photographic plates, concluded that low-mass s-elements (Sr, Y, Zr) are in nearly solar proportions in J-stars, while rare-earth elements are overabundant by factors of 10-100. However, a further revision by Utsumi (1985) showed that in J-type stars, abundances of s-process elements with respect to Fe are nearly normal. Kilston (1975) derived Zr/Ti ratios in two J-stars (Y CVn and RX Peg) and Dominy (1985), also using low dispersion spectrograms, defined an abundance index by comparing the intensity of some lines of low-mass s-process elements with those of metals. Dominy drew similar conclusions to Utsumi with the possible exception of the star WZ Cas. Recently, Lambert et al. (1986) and Harris et al. (1987) have focussed their attention on the CNO content of J-stars. The study of correlations between different chemical species in J-stars might cast more light upon their origin and evolution. In this work we perform a detailed abundance analysis of twelve galactic J-type carbon stars using very high resolution and signal-to-noise ratio echelle spectra from the visible to the near infrared. To do this we make use of up-to-date model atmospheres, and atomic and molecular data. We focus our attention mainly on the determination of s-process element abundances. Our results, together with CNO and Li abundances determined in other studies, are contrasted with theoretical stellar models to find an evolutionary status for J-type stars. ## 2 Observations The observations were made during 1997 and 1998 at two different observatories. At the Calar Alto observatory we used the 2.2 m telescope in October-97 and July-98 during three nights in each month. For these observations a fibre optics Cassegrain Echelle Spectrograph was used (FOCES; Pfeiffer et al. 1998). This instrument uses a 1024$`\times `$1024 CCD with 24 $`\mu `$m pixel size. The FOCES image covers the visible spectral region from 0.38 to 0.96 $`\mu `$m in about 80 orders with full spectral coverage. The resolving power achieved was $`40000`$ with a two pixel resolution element. At La Palma observatory the 4.2 m WHT was used with the Utrech Echelle Spectrograph during one night in December-97 and May-98. The spectral range coverage with this instrument, using the 79.0 lines/mm grating, is from 0.43 to 1.0 $`\mu `$m with some gaps between orders. The projected size of the slit in the 2048$`\times `$2048 CCD was around two pixels (24 $`\mu `$m) which gave a resolving power of $`50000`$. We observed a total of 43 galactic carbon stars selected from the two-micron sky survey by Claussen et al. (1987). Twelve J-type carbon stars of this sample are analyzed here (see Table 1). The analysis of the additional 31 normal (N-type) carbon stars will be presented in a future work. We used standard IRAF packages and procedures to perform bias level, dark current and scattered light subtraction and to prepare a normalized flat-field image to remove pixel-to-pixel sensitivity fluctuations. A Th-Ar comparison lamp gave enough lines in all the echelle orders to perform an accurate wavelength calibration. We carefully identified non-saturated Th-Ar emission lines and adjusted third-fourth order polynomials to obtain a calibration fit better than 10 mÅ in the residuals. The calibrated spectra were then divided by the spectrum of a hot, rapidly rotating star located in the sky as close as possible to the target star to eliminate telluric absorptions. We note however, that most of the lines used in the abundance analysis are not affected by terrestrial features (see below). Finally, different images of the same object were co-added after extraction and calibration to obtain the final spectrum. The S/N ratio achieved in the final spectra varies from the blue to the red orders. At $`4500`$ Å the S/N is 40-50 while at $`8000`$ Å the S/N frequently exceeded 400. In contrast, below $`4400`$ Å the S/N ratios are poor because at these wavelengths carbon stars (J and N) are very difficult to observe. The reason for this flux depression is still a matter of controversy. In J-stars, particularly rich in <sup>13</sup>C, the absorption of triatomic carbon compounds can become nearly continuous at the shorter wavelengths. On the other hand, Johnson et al. (1988) have pointed out that at the low temperatures of these stars, the resonance lines of some atoms become so wide that they depress broad areas of the continuum. This makes it almost impossible to observe the interesting Tc I resonance lines at $`4250`$ Å in C-stars. ### 2.1 Individual stars The spectral types shown in Table 1 are based on the C-system classification by Keenan & Morgan (1941) revised by Yamashita et al. (1977). Our stars fullfil the criteria for the <sup>13</sup>C abundance by Keenan (1993) to be classified as J-stars: ratio of <sup>13</sup>C<sup>12</sup>C $`\lambda 4744`$ Å to <sup>12</sup>C<sup>12</sup>C $`\lambda 4737`$ Å in blue: ratios of <sup>12</sup>C<sup>12</sup>C $`\lambda 6191`$ Å to <sup>13</sup>C<sup>12</sup>C $`\lambda 6168`$ Å, <sup>13</sup>C<sup>12</sup>C $`\lambda 6102`$ Å to <sup>12</sup>C<sup>12</sup>C $`\lambda 6122`$ Å and <sup>13</sup>C<sup>14</sup>N $`\lambda 6260`$ Å to <sup>12</sup>C<sup>14</sup>N $`\lambda 6206`$ Å. However, for some of them the spectrum is misleading, which might produce to a problematic spectral classification. For instance, WZ Cas shows weaker CN and C<sub>2</sub> band absorptions than the other J-stars in the sample. It also shows very strong Na D lines and its spectrum does not look as crowded as the rest of the J-stars. In fact, atomic lines in WZ Cas are more easily identified. Indeed, this happens when the C/O ratio in the atmosphere is very close to unity, a characteristic which defines a SC-type carbon star. Since our C/O estimate in this star is $`1.01`$, we believe that WZ Cas has to be considered a SC-star rather than a typical J-star. Furthermore, it is the most luminous star in our sample (M$`{}_{\mathrm{bol}}{}^{}6.44`$), which could indicate that WZ Cas belongs to a different population (massive) of C-stars or that it is in a different evolutionary status (more evolved) than the rest of the J-stars in our sample. However Lorenz-Martins (1996) argues that it is difficult to separate J-type stars and SC stars only by spectroscopy results. Modeling the dusty envelope of this star he concludes that WZ Cas could be a J-type carbon star. Less obviously, the same peculiarities are observed in the spectrum of WX Cyg. In fact, Ohnaka & Tsuji (1996) considered this star to be SC type although, as far as we know, no other work in the literature classifies this star as SC-type. For this star we do not have a clear opinion. On the other hand, FO Ser is also classified in the literature as a late R6 type carbon star (Stephenson, 1989). We consider it a typical J-star but we note that it presents H<sub>α</sub> in absorption which is not seen in other J stars or in N or late R stars. ### 2.2 Luminosities of the stars To estimate absolute bolometric magnitudes M<sub>bol</sub> of our stars (Table 1) we have used the empirical relationships between M<sub>bol</sub> and M<sub>K</sub>-M<sub>V</sub> for C-stars obtained by Alksnis et al. (1998). Infrared absolute magnitudes M<sub>K</sub> are frequently used in connection with distance determinations where it is assumed that all AGB stars have the same M$`{}_{\mathrm{K}}{}^{}8.1`$ (Frogel et al., 1980). To obtain M<sub>K</sub> and M<sub>V</sub>, the K and V average values (the stars studied are variable!) quoted in the SIMBAD database were used. K and V magnitudes were corrected for interstellar extinction according to the galactic extinction model by Arenou et al. (1992). Distances were derived from parallax measurements by HIPPARCOS. Note that some parallaxes have considerable errors (see Table 1); for instance, the large uncertainty in the parallax for UV Cam produces an unlikely value for M<sub>bol</sub> in this star. Thus, the bolometric magnitudes in Table 1 have to be considered as average values and only indicative. ## 3 Analysis ### 3.1 Effective temperatures For the majority of the stars studied here the effective temperature is derived by Ohnaka & Tsuji (1996,1999). These authors derived effective temperatures for C-stars using the infrared flux method. The $`L^{}`$band (3.7 $`\mu `$m) is used as infrared flux and T<sub>eff</sub> values are obtained from the calibration log$`(f_{Bol}/f_L^{})`$ vs. T<sub>eff</sub>, where $`f_{Bol}/f_L^{}`$ is the ratio of observed bolometric flux to infrared flux. The T<sub>eff</sub> values in Table 2 marked with an asterisk() are derived in this way. For the rest of the stars we used the $`(JL^{})_o`$ vs. T<sub>eff</sub> calibration also described by Ohnaka & Tsuji (1996). Infrared photometry for our stars is taken from Noguchi et al. (1981) and Fouqué et al. (1992). Effective temperatures for J-stars derived in this way do not differ significantly from those derived in N- and SC-type carbon stars. The estimated error in T<sub>eff</sub> is $`\pm 150`$ K (see Ohnaka & Tsuji 1996, for details). ### 3.2 Model atmospheres The set of models used in this analysis was computed by the Uppsala group (see Eriksson et al. 1984, for details). The models cover the T$`{}_{\mathrm{eff}}{}^{}=25003500`$ K, C/O$`=1.01.35`$ ranges and all have the same gravity log g$`=0.0`$. The input elemental abundances adopted for the J-star models were the solar values, with the exception of C, N and O which were assumed to be altered relative to the Sun. The CNO abundances in the model atmosphere for a given star were taken from the literature (Lambert et al., 1986; Abia & Isern, 1997). For each star a model atmosphere was interpolated in T<sub>eff</sub> and C/O ratio in this grid. A typical microturbulence velocity for AGB stars $`\xi =3`$ kms<sup>-1</sup> was adopted or taken from the literature when available (Lambert et al., 1986). Table 2 shows the atmosphere parameters used in the analysis. ### 3.3 Identification of the atomic lines The greatest difficulty in the analysis of atomic lines of carbon stars lies in the heavy blend effect of molecular bands. The problem of blending is even more serious in J-stars because of the strong C<sub>2</sub> and CN band absorptions. For this reason, identification of spectral lines in carbon stars is very difficult and, moreover, detailed laboratory data of these molecules are usually scarce. Utsumi (1967) carried out line identification in two carbon stars in the visual region and found that the blend effect of molecular bands is not so serious in the region between $`\lambda 44004500`$ and $`\lambda 47504900`$ Å. One difficulty arises, however, because at these wavelength ranges carbon stars (particularly of J-type) are very faint, so that very long time exposures are needed to obtain good S/N spectra. Wallerstein (1989) (7 Åmm<sup>-1</sup>) and Barnbaum et al. (1994) ($`3`$ Åmm<sup>-1</sup>) identified atomic lines in SC-and C-stars, respectively, in a more crowded region ($`\lambda 50008000`$ Å). The accuracy in their line identifications ranges from 0.03 to 0.08 Å depending on the line intensity. Basically, we have used these three atlases for atomic line identification. As a first step, we corrected the spectra for the stellar radial velocity using the wavelengths of some easily recognized features: the two Na D lines, the Li I resonance line at $`\lambda 6708`$ Å and the K I line at $`\lambda 7698`$ Å, together with a few intense Fe I and Ca I lines. The standard deviation around the mean Doppler shift obtained from these features was less than $`\pm 0.1`$ Å. This was the maximum wavelength shift allowed between measured and expected wavelengths to consider a feature as a good identification. Next, we followed the same criteria as in Abia & Wallerstein (1998) (hereafter Paper I) to consider an identification as useful for abundance analysis, namely: first, the line should not present a clear visual blend with any adjacent line and second, the local continuum around the line should be reasonably placed within $`5\%`$ of uncertainty. Of course, in no way is it possible to establish the real continuum in our stars. We can only define a local or pseudo-continuum connecting the highest flux points near the line of interest. We also tried to use the weakest possible lines, log $`[W(\lambda )/\lambda ]4`$, although in most of the stars this was not possible. When applying these criteria we immediately realized that even our spectral resolution is not enough to clearly resolve most of the atomic features in the spectra. In fact, the majority of the atomic identifications in Utsumi’s list are clear blends. Therefore, most of the equivalent width measurements by Utsumi (1970) are overestimations, and the corresponding atomic line is not useful for analysis. This explains the s-process element overabundances derived by this author. The atomic identification list by Wallerstein (1989) and Barnbaum et al. (1994) covers a very crowded region in J-stars and only a few of the atomic lines there were considered good candidates in our stars. In this way we drew up a first atomic line list. Finally, each line was checked for possible atomic and/or molecular contributions not clearly seen in the spectra as a blend. When any feature was suspected of contributing to the atomic line selected, we estimated its contribution by computing theoretical equivalent widths using the model atmosphere parameters shown in Table 2. If the estimated contribution was to be higher than $`15\%`$ of the total equivalent width measured, the line was ruled out for analysis. The final atomic line list is shown in Table 3. As can be seen, very few lines were found to be useful for analysis. Note that several hundred atomic lines were searched in each star using the above selection criteria. For some species only one line was found. Table 3 also shows the total equivalent widths measured. To do this we used the SPLOT program of the IRAF package. In some cases only upper limits are derived since, even when the line appears to be free of blends, we believe the uncertainty in the location of the continuum near the line could be higher than 5%. We estimate the error in the equivalent width from the theoretical expression given by Cayrel (1988) (see also Paper I). The uncertainty ranges from $`\mathrm{\Delta }\mathrm{W}(\lambda )=10`$ to 35 mÅ, according to the line intensity and to the S/N of the spectrum, with the main uncertainty being introduced by the continuum placing. When possible, gf values were derived from identification and equivalent width measurements in the Solar Atlas by Moore et al. (1966), using solar abundances from Anders & Grevesse (1989). We used the solar model atmosphere by Holweger & Müller (1974) with parameters T$`{}_{\mathrm{eff}}{}^{}=5780`$ K, log g$`=4.44`$ and microturbulence variable with optical depth. Note that the possibility of using the Sun as a standard to derive astrophysical gf-values is greatly limited by the weakness of the lines used here in the solar spectrum. Specific references for individual lines are given in Table 3, otherwise we used the gf-values given in the VALD database (Piskunov et al., 1995). Finally, we considered the hyperfine structure using such information as was available: McWilliam (1998) for Ba: for Tc and Rb lines see the discussion in Paper I and references therein. Broadening by radiation damping was calculated as in Edvardsson et al. (1993), when not given explicitly by VALD. Finally, the classical van der Waals damping constant of the atomic lines was modified, also following Edvardsson et al. ## 4 Abundance results ### 4.1 Lithium All our stars show a strong $`\lambda 6708`$ Å Li I absorption (see Figure 1). In WZ Cas and WX Cyg the equivalent width of this spectral feature is certainly larger than 1 Å. Abia, Pavlenko & de Laverny (1999) discuss the formation of Li I lines in C-rich atmospheres considering also N-LTE effects. They conclude that among the Li lines available for analysis in AGB stars ($`\lambda 4603`$, $`\lambda 6104`$, $`\lambda 6708`$ and $`\lambda 8126`$ Å), the subordinate line at $`\lambda 8126`$ Å is probably the most reliable, on the basis of high S/N spectra and spectral synthesis. This line forms deep enough in the atmosphere where the uncertainties in the model atmosphere structure of AGB stars are smaller. N-LTE effects for this line are also weak ($`0.2`$ dex) and the continuous opacity coefficient seems to be well reproduced by model atmospheres in this wavelength range. Unfortunately, in our stars, except WZ Cas and WX Cyg, the $`\lambda 8126`$ Å Li line is weak and very crowded with strong CN absorptions in such a way that synthetic fits to this line differing by $`0.20.3`$ dex in the Li abundance do not significantly alter the quality of the fit. Thus, we decided to use the $`\lambda 6708`$ Å line, which is much more sensitive to abundance variations, except for the stars WX Cyg and WZ Cas. For these two stars we used instead the $`\lambda 8126`$ Å Li I line. In any case, all the Li abundances were derived by spectrum synthesis and corrected by N-LTE effects according to Abia, Pavlenko & de Laverny (1999) (see this paper for details). However, one has to be very careful when interpreting the Li abundances derived from the resonance Li I line. The presence of a circumstellar component in the $`\lambda 6708`$ Li absorption would lead us to overestimate the Li abundance when derived from this line. In fact, VX And, BM Gem and V614 Mon show weak blueshifted Na D line absorptions, probably indicating the presence of a circumstellar envelope in these stars. This circumstellar absorption is, however, not observed in the strong $`\lambda 7698`$ Å K I resonance line which should form at about the same depth in the atmosphere than the Li resonance line (see Barnbaum 1992). The lack of K I circumstellar absorption probably rules out significant contamination of the photospheric feature. Final Li abundances are shown in Table 4 in the scale log $`ϵ`$(Li)$`=12+`$ log(Li/H), where Li/H is the abundance of Li by number. From Table 4 it is clear that all the stars have unusual Li abundances (log $`ϵ`$(Li)$`1`$), larger than the typical value found in normal C-stars (log $`ϵ`$(Li)$`0.0`$), but significantly smaller than those found in the so-called super Li-rich. WX Cyg and WZ Cas are certainly super Li-rich stars, although these stars may not be J-type stars (see above). The formal uncertainty in Li abundances of Table 4 ranges from 0.3-0.4 dex. Figure 2 shows the correlation of Li abundances versus <sup>12</sup>C/<sup>13</sup>C ratios found in J- (this work) and N-type carbon stars (Abia & Isern, 1997). J-type stars are all Li-rich and note that there are also some Li-rich N-type stars. The formal error in the <sup>12</sup>C/<sup>13</sup>C ratios in Figure 2 is $`\pm 6`$ (see Abia & Isern 1997). ### 4.2 Technecium The presence of Tc in the atmosphere of AGB stars (in fact <sup>99</sup>Tc) is commonly interpreted as evidence of the operation of the s-process within stars. This study is the first detailed search for the presence of this element in J-type stars. As mentioned above, the three resonance Tc I lines near $`\lambda 4250`$ Å are unaccessible in C-stars because of the strong flux depression in these stars below $`\lambda 4400`$ Å. Therefore we have used, as in Paper I, the intercombination and weaker Tc line at $`\lambda 5924.47`$ Å. We followed the same procedure in the analysis as in Paper I. The reader is referred to this paper for a detailed discussion of the identification of the Tc blend and the choice of the gf-value and the hyperfine structure of this line. As in Paper I, the $`\lambda 5924`$ Tc blend is not well reproduced by synthetic spectra, in particular the red-wing of the line. Thus, although in some stars the Tc blend was apparently well reproduced, we prefer to be cautious and record the Tc abundance as equal-to-or-less-than. Figure 3 shows a clear example of this in the star WZ Cas. Tc may be present in this star but the poor fit to the red part of the line prevents us from asserting a definitive detection. We place an upper limit of log $`ϵ`$(Tc)$`1`$. For WX Cyg, the other possible detection, we set log $`ϵ`$(Tc)$`0.7`$. Note that these two possible detections are of the same level as the Tc upper limits set for the sample of SC stars analyzed in Paper I. In the remaining stars, the best fit to the Tc blend is compatible with no-Tc, i.e. a synthetic spectrum with no-Tc does not differ from another one computed with a small Tc abundance. For these stars we quote a no entry in Table 4, meaning that Tc, very probably, is not present. An example of this is shown in Figure 3 for the star UV Cam. In fact, for this star, Y CVn and RX Peg we were able to obtain spectra in the $`\lambda 4250`$ Å region with a high enough S/N ratio to check for the presence of the Tc resonance lines. We have not quantitatively analyzed these spectra because we lack the appropriate atomic and molecular lines in this spectral region, but a careful search for the resonance lines confirmed the absence of technetium in these three stars. We agree with the previous finding by Little, Little-Marenin & Hagen (1987) in Y CVn. Barnbaum et al. (1991) report the possible detection of Tc in two J-stars, EU And and BM Gem (star studied here). Their argument is based on the presence of a strong absorption at $`\lambda 6085`$ Å which is partially due to the $`\lambda 6085.22`$ Å Tc I line. From these authors, the fact that the $`\lambda 6085`$ absorption appears which similar intensity only in those carbon stars where Tc has been detected unambiguously using the blue lines, supports the identification of the feature at $`\lambda 6085.22`$ Å as Tc. Our quantitative analysis of the $`\lambda 5924`$ Tc feature in BM Gem is, however, compatible with a non-detection. We note that the $`\lambda 5924`$ Tc feature is a factor $`2`$ more intense than the $`\lambda 6085`$ one (Garstang, 1981), therefore the presence of Tc in BM Gem in any measurable amount should have appeared in our analysis. Furthermore, the $`\lambda 6085`$ Tc feature is strongly blended with a Ti I line of moderate intensity ($`\chi =1.05`$ eV, log gf$`=1.35`$) and some CN and C<sub>2</sub> absorptions, which necessarily requires a spectral synthesis analysis to confirm the detection. Leaving apart the possible presence of Tc in BM Gem for a further and accurate analysis and the upper limits set for WZ Cas and WX Cyg, possible SC-type stars, we can conclude that most of J-stars do not show Tc. ### 4.3 Rubidium The Rb abundance is a monitor of the neutron density at which the s-process operates in AGB stars. Therefore the derivation of Rb abundances in these stars is extremely important, specifically the abundance ratios between Rb and its neighbors in the periodic table (Zr, Sr). We have used the resonance line at $`\lambda 7800.23`$ Å to derive Rb abundances. The other accessible line at $`\lambda 7947`$ Å is much weaker and very crowded with CN lines in cool stars. Nevertheless, the resonance line is also blended, and so Rb abundances have to be derived from spectral synthesis. We have used the same atomic and molecular line list in the Rb spectral region as in Paper I with the addition of some C<sub>2</sub> lines (including the <sup>13</sup>C isotope) computed by P. de Laverny (private communication). We refer again to Paper I for a discussion of the identification of the atomic and molecular lines contributing to the $`\lambda 7800`$ Å blend. The Rb line is represented by the hyperfine structure components of both isotopes <sup>85</sup>Rb and <sup>87</sup>Rb in a terrestrial ratio (<sup>85</sup>Rb/<sup>87</sup>Rb=2.59) with gf-values taken from Wiese & Martin (1980). Only in three stars (WX Cyg, WZ Cas and V353 Cas) does the Rb line appear clearly as a prominent absorption in the background of CN lines. In the remaining stars the Rb line is not distinguished from the background of lines (see Figure 4). Table 4 shows the Rb abundance derived in our stars relative to their mean metallicity \[M/H\]<sup>1</sup><sup>1</sup>1We adopt here the usual notation \[X\]$``$log(X)-log(X) for any abundance quantity X.. We adopt the solar photospheric Rb abundance by Anders & Grevesse (1989): log $`ϵ`$(Rb)=2.60$`\pm 0.15`$. If the meteoritic Rb abundance is preferred, the \[Rb/M\] values in Table 4 have to be increased by 0.2 dex. From Table 4 it is apparent that the \[Rb/M\] ratios derived are remarkably low. For some stars the best fit is compatible with no Rb. Nevertheless, we believe that our Rb abundances could be, and in some cases are, lower limits. The are several reasons for this: first, some metallic lines (Fe,Ni) near the Rb line are best fitted by theoretical spectra assuming abundance values which are systematically lower by $`0.20.3`$ dex than that of the mean metallicity of the star derived from other metallic lines (see Table 3). Thus, the \[Rb/M\] ratios should be increased by this factor if \[M/H\] is derived from the atomic lines near Rb. On the other hand, our synthetic spectra typically give a very strong $`\lambda 7800`$ Å Rb I absorption even for very low Rb abundances. We found the same figure when deriving abundances from other resonance or very low excitation energy lines of elements with a similar ionization potential to Rb (4.18 eV). Consider, for instance, the resonance $`\lambda 6708`$ Å Li I (5.39 eV) and $`\lambda 7698`$ Å K I (4.34 eV) lines, two alkali elements with a similar atomic structure. In fact, synthetic fits to the K I line give low potassium abundances (undersolar). No nuclear mechanism able to destroy potassium in stars is known. This effect with K however, was not found by Plez, Smith & Lambert (1993) in M giants of the Magellanic Clouds. Non-LTE effects (overionization) could be the reason for this as in the case of the strong resonance Li I line in some C-stars (Abia, Pavlenko & de Laverny, 1999). N-LTE corrections on this line might extend to until +0.6 dex in the sense of N-LTE minus LTE abundances. Similar unexpected low \[Rb/M\] ratios were also derived by Plez, Smith & Lambert (1993) and Lambert et al. (1995). However, these authors conclude that NLTE effects on Rb are probably weak, from analysis of the Rb line in Beltegeuse and Aldebaran, two galactic M supergiants whose atmospheres are presumed to be similar to those of the O-rich AGB stars they studied. A quantitative N-LTE study of the formation of the Rb line in cool C-rich atmospheres is needed before this question can be answered. Typical maximum uncertainties in the atmosphere parameters ($`\mathrm{\Delta }\mathrm{T}_{\mathrm{eff}}=\pm 200`$ K; $`\mathrm{\Delta }\xi =\pm 1\mathrm{kms}^1`$; $`\mathrm{\Delta }\mathrm{CNO}/\mathrm{H}=\pm 0.3`$ dex; $`\mathrm{\Delta }\mathrm{C}/\mathrm{O}=\pm 0.05`$ dex) and $`5\%`$ in the continuum added quadratically, represent a maximum uncertainty of $`0.4`$ dex in the absolute abundance of Rb derived. The formal uncertainty concerning the \[Rb/M\] value is probably less than this, because some of these sources of uncertainty affect the \[M/H\] values in a similar way. Thus, when deriving the \[Rb/M\] ratio, many errors would be canceled out. However, we avoid to estimate neutron densities in the s-process site from the derived Rb/Zr or Rb/Sr ratios due to the uncertain Rb abundances as mentioned above. ### 4.4 Metallicity and Heavy Elements The abundances of metals were derived from the usual method of equivalent width measurements and curves of growth calculated in LTE. Ca, V, Fe and Ti abundances were used as a measure of the metallicity of the stars. The \[M/H\] value shown in Table 4 is the mean metallicity obtained from these elements. The upper limits in Table 3 were not considered when deriving \[M/H\]. In the star Y CVn we were not able to identify any metallic line useful for abundance analysis. For this purpose, we adopted the metallicity obtained by Lambert et al. (1986) from several Fe and Ca lines. Note that the number of metallic lines analyzed per star is rather low: minimum, one and maximum, eight for WZ Cas. This star is the only one for which a reasonable statistic with Fe lines (five) can be performed. We found a mean dispersion of $`\pm 0.1`$ dex around the mean iron abundance derived, which is compatible with the error introduced by the uncertainty in the equivalent width measurements. On the other hand, the elements having isotopes formed by neutron captures have very few useful lines in the visible spectra of J-stars. Note the significant number of empty entries or upper limits in Table 3. WZ Cas is again the sole star where it is possible to detect a significant number of heavy element lines. A resolving power of $`10^5`$ is needed to perform an accurate analysis of these stars. This means that abundance analyses in C-stars based on intermediate-low resolution spectra and/or on the visual intensity of spectral lines can lead to important errors. For example, the $`\lambda 4607.34`$ Å Sr I, $`\lambda 4554.04`$ Å Ba II and $`\lambda 6709.49`$ Å La I features, used by Dominy (1985) to define an abundance index of these s-process elements, appeared in our spectra as very crowded blends. At these wavelengths many CN and C<sub>2</sub> lines contribute significantly in C-stars. Therefore, a high intensity of such lines does not necessarily mean an enhancement of Sr, Ba or La. This kind of analysis is only useful in relative abundance studies between stars, not to derive absolute abundances. Figure 5 compares the strength of the Ba II line (4934.07 Å) found in three J-type stars. To establish a wider comparison, the spectrum of a normal (N) C-star (Z Psc), presumed to be rich in Ba (Utsumi, 1985), is included. All the spectra in Figure 5 were obtained in the same way with echelle spectrographs. The Ba II line is clearly strong in Z Psc, reflecting its probable Ba overabundance, but among the J-stars this line is not so intense. Since in general few lines are available, it is not practical to examine the distribution of abundance against atomic number or even the mean difference between low-mass (Sr-Y-Zr) and high-mass (Ba-La-Ce-Nd-Sm) s-elements star by star. Instead, we derive the mean heavy-element enhancement \[$`<h>`$/M\] shown in Table 4. To derive this we did not consider upper limits or the uncertain Rb abundances. As in Paper I, an estimate was made of the theoretical errors concerning the derived metal abundances due to the uncertainties in the atmosphere parameters of the stars. The formal error due to errors in T<sub>eff</sub>, microturbulence, CNO abundances, the C/O ratio in the atmosphere, equivalent width measurements and the location of the continuum, added quadratically, is $`\pm 0.30.6`$ dex, depending mainly on the intensity and excitation energy of the line as well as on the ionization state of the line considered. The microturbulence parameter is also an important source of uncertainty since most of the abundances are derived from strong lines, in the flat part of the curve of growth. For example, a variation of $`\mathrm{\Delta }\xi =\pm 1`$ kms<sup>-1</sup> produces a change of $`\pm 0.4`$ dex in the barium abundance derived from the strong Ba II line at $`\lambda 4934`$ Å. Unfortunately, given the large uncertainties in the equivalent width measurements and the few number of lines identified (§3.3), it was impossible to estimate $`\xi `$ using the requirement that individual abundances derived from lines of different intensity have to be nearly equal. Taking the error bar above into account our results in Table 4 show that J-type C-stars are of near solar metallicity $`\overline{[\mathrm{M}/\mathrm{H}]}=0.12\pm 0.16`$ and do not show the sizable heavy element enhancements typical of S or SC stars (Smith & Lambert, 1990; Abia & Wallerstein, 1998). The mean heavy element enhancement among the J-stars in the sample is \[$`<h>`$/M\]$`=0.13\pm 0.12`$, which is compatible with non-enrichment. Considering individual stars (see Table 4), in some of them there is a hint of a heavy-element enrichment, but given the small number of lines analyzed and the large errors, this has to be considered with caution. Comparison with the results obtained by Utsumi (1985) for the stars in common (UV Cam, WZ Cas, Y CVn and RY Dra) is difficult because of the different methods used in the analysis. Furthermore, Utsumi uses only Ti as the metallicity indicator and refers all the abundances to this element. Instead, we have used Fe, Ca, V and Ti to obtain the metallicity \[M/H\]. However, the abundance ratios found here agree with those of Utsumi in the stars in common between the error bars (Utsumi estimated an accuracy of about 0.4 dex). ## 5 Evolutionary Considerations J-type carbon stars are not rare, amounting to $`15\%`$ of C-rich giants, and therefore they should represent a stage of evolution that is available to a significant fraction of stars, and are not the result of anomalous initial conditions or statistically unlikely events. The chemical abundances found in the present and other studies offer constraints to certain scenarios that have been offered to account for the existence of these C-stars. In the following paragraphs, we use these abundance results to discuss the evolutionary status suggested and to propose new ones. First, let us recall the chemical properties of J-stars: a) they are certainly carbon stars (C$`>`$O) and show very low carbon isotope ratios ($`<15`$). In many of them the <sup>12</sup>C/<sup>13</sup>C ratio is equal to the CNO cycle equilibrium value. b) An important fraction ($`75\%`$) have enhanced Li (log $`ϵ`$(Li)$`>1`$) although the majority are Li-rich rather than super Li-rich objects. c) They are solar metallicity stars. d) They do not show Tc or s-process element enhancements in their atmospheres. Obviously, exceptions to these figures can be found but we are only discussing these stars on the basis of their most common properties. For instance, the carbon star UV Aur (a symbiotic star) is classified as J-type, although it shows Tc (Little, Little-Marenin & Hagen, 1987) and does not present Li (Boffin et al., 1993). Unfortunately, we could not include this star in the present study. Figure 6 shows the position of our J-stars in an observational H-R diagram, including some galactic R-type and N-type carbon stars with absolute magnitudes also derived from the HIPPARCOS parallaxes (see Alsknis et al. 1998). From this figure, one might consider J-stars as transition objects between R-stars and N-stars. This is reinforced considering the fact that most J-stars are irregular or semi-irregular variables (very few Miras are found among them) with not very large pulsation periods, which is a characteristic of the less evolved carbon stars. Furthermore, on average the envelopes of J-stars are thinner than those of ordinary carbon stars (Lorenz-Martins, 1996), which could also suggest that this class of objects is in the very early stages of carbon star evolution. Current AGB models (Sackmann & Boothroyd, 1992; Lattanzio, 1999; Blöcker, 1998) can obtain C-rich (C$`>`$O) envelopes and low carbon isotope ratios in stars with initial mass M$`4`$ M through the successive He-shell flashes and TDU episodes coupled with the operation of HBB at the base of the convective envelope. These stars can also be, for a long period of time, Li-rich AGB stars with peak Li abundances in the range log $`ϵ`$(Li)$`34`$. However, the operation of HBB leads to the transformation of <sup>12</sup>C into <sup>14</sup>N; thus nitrogen is expected to be enhanced in the envelope of these stars. The nitrogen abundances derived in some J-stars (Lambert et al. 1986) show a normal N/O ratio, much lower than that expected on the basis of the CNO cycle operation in HBB. The <sup>17</sup>O/<sup>16</sup>O and <sup>18</sup>O/<sup>16</sup>O ratios measured in J-stars (although with an important error bar) also argue against a pure CNO cycle interpretation of the J-stars chemical anomalies (Harris et al., 1987). Models by Lattanzio & Forestini (1998) can only obtain a C-rich, Li-rich, <sup>12</sup>C/<sup>13</sup>C low and N/O$`<1`$ AGB star in a very narrow range of stellar masses (M$`5`$ M), with a specific metallicity (Z$``$ Z/3) and for a very short period of time ($`10^4`$ yr). In this context, the number of J-stars expected would be very low, which is in contrast with the significant number observed. On the other hand, these objects would be fairly luminous (M$`{}_{\mathrm{bol}}{}^{}<6`$), and should present some s-process element enhancement (at least of low-mass Sr-Zr-Y, see Vaglio et al. 1998). None of this is observed (see Tables 1 and 4). Thus, standard AGB models for masses M$`4`$ M are very difficult to reconcile with the observed properties of J-stars. Note in addition that there is observational evidence indicating a low-mass (M$`23`$ M) progenitor star for most of the J-stars studied here (e.g. Claussen et al. 1987). Most of our stars are not very luminous objects (M$`{}_{\mathrm{bol}}{}^{}>5`$). Their luminosity is of the order of the predicted value for low-mass stars during or near the AGB phase. Although the formation of low-mass C-stars has recently been found to be possible (Straniero et al., 1997), HBB is not found in low-mass AGB models because of the low temperatures reached at the base of the convective envelope. One has to advocate, therefore, the existence of a non-standard mixing mechanism which transports material from the bottom of the convective envelope into deeper and hotter regions (basically the H burning shell) where cool processing might occur. This hypothetical mixing mechanism, perhaps induced by meridional currents, was proposed by Wasserburg, Boothroyd & Sackmann (1995) and has been shown to reproduce the CNO isotope anomalies found in some low-mass red giants (Boothroyd & Sackmann 1999). Under certain conditions, it can also create <sup>7</sup>Li via the Cameron and Fowler mechanism (Cameron & Fowler, 1971), thus accounting for the recent discovery of surprisingly high lithium abundances in some low-mass red giants (Brown et al., 1989; Wallerstein & Sneden, 1982; de la Reza et al., 1997). Boothroyd & Sackmann (1999) suggest that this extra-mixing and cool bottom processing could also occur in low-mass AGB stars and account for the <sup>18</sup>O depletion and low <sup>12</sup>C/<sup>13</sup>C ratios found in the J-stars analyzed by Harris et al. (1987). The operation of this mechanism on the early-AGB or just after the onset of the helium shell flashes is preferred since cool bottom processing appears to become weaker as the star ascends the AGB phase. This point might be compatible with the suggestion (see Figure 6) that J-stars are not very evolved AGB stars but are just on the verge of becoming normal N-type carbon stars. In that case, little or no s-process element enhancement would be expected, as a significant number of TDU episodes are needed. This might also be compatible with the abundance results presented here. A quantitative analysis of the operation of this extra-mixing and cool bottom processing and the subsequent envelope chemist on the AGB is currently in progress (Sackmann & Boothroyd 1999, in preparation). With the above AGB scenarios encountering difficulties in explaining the existence of J-stars, we examine a scenario outside the AGB phase: the mixing at the He core flash. This mechanism has already been proposed to explain the evolutionary status of the R-stars (Dominy 1985). Note that as far as the chemical composition is concerned, R-stars and J-stars only differ in the presence of Li and slightly lower <sup>12</sup>C/<sup>13</sup>C ratios in the latter. Thus, there is the suspicion that R-stars may be the ancestors of J-stars (see Figure 6) after an additional mixing event (second dredge-up?). Very few studies of the core He-flash have been published. In an off-centre core flash hydrodymanic calculation, Deupree & Cole (1981, 1983) show that a bubble of low density and high temperature (T$`4\times 10^8`$ K) can form and mix the He-core and H-shell matter. If in some red giant stars the core flash does introduce C-rich material into the H-burning shell, there is reason to believe that the results of He-burning and CN-processing may reach the convective envelope and ultimately the surface. The main question is whether the He-flash is able to produce and mix enough <sup>12</sup>C into the envelope to transform the star into a C-star. Investigations by Mengel & Gross (1976) of non-central flashes (in fact, near the core boundary) within a rapidly rotating core show that a series of flashes could occur and build up the C abundance in the envelope through successive flash and mixing events. Recently, Deupree & Wallace (1996) have re-examined the He-core flash performing hellium flash calculations of different intensity. The violence of the flash is mainly governed by the degree of degeneracy where the explosion occurs. The authors estimate the surface abundance anomalies produced by the different He-flashes. They show that the primary material mixed into and above the hydrogen shell in all cases is <sup>12</sup>C. The other major products are the result from hot $`\alpha `$ captures that occur during the flash (<sup>20</sup>Ne, <sup>24</sup>Mg, <sup>28</sup>Si and <sup>32</sup>S) and, if the hydrogen shell is penetrated at reasonably high temperature, some <sup>14</sup>N. Observable enhancements of <sup>12</sup>C in the envelope are favoured in very metal poor, low-mass envelope red giants. For moderate peak temperatures of the flash ($`9\times 10^8`$ K) important <sup>12</sup>C enhancements (by a factor of 70) can also be obtained in solar metallicity stars, in such a way that the star might become a carbon star. However, in order to account for the observed chemical peculiarities of J-stars, fine tuning of the models is required. First, on the way, <sup>12</sup>C must be exposed to protons and much of it be converted to <sup>13</sup>C to get a low <sup>12</sup>C/<sup>13</sup>C ratio in the envelope. In addition, to prevent the release of neutrons via the <sup>13</sup>C($`\alpha `$,n) reaction and so the creation of s-process nuclei, the temperature in the mixing zone must not exceed $`10^8`$ K (interestingly, Deupree & Wallace claim that their flash computations do not produce s-process elements). Finally, Li production would require temperatures not exceeding $`5\times 10^7`$ K in the processing zone. At least the last temperature requirement appear rather difficult to attain (Lattanzio, private communication). Perhaps, Li can be produced after the He-flash by an additional mixing event between the convective envelope and the H-shell. Obviously, the He-flash scenario merits further hydrodynamic (3-D) studies. Note, the He-flash occurs in low-mass and low-luminosity objects such the stars studied here. Finally, we consider the mass-transfer scenario in a binary system. The existence of S stars with no Tc, as predicted by the mass-transfer paradigm, is now well established (Jorissen, Frayer and Johnson, 1993). Whether this scenario can also be applied to C-stars is not yet firmly demonstrated, although Barnbaum (1993) found 16 Tc-poor stars in a sample of 78 C-N stars with Ba excess. However, it is difficult to explain the absence of s-process element enhancement and the C/O ratios in our stars within this scenario. In principle, the accreted material must be extremely carbon-rich; the donor star should be a normal C-star with probably enhanced s-nuclei in the envelope. Dominy (1985) estimated a C/O$`5`$ in the material accreted by a typical $`1`$ M red giant when applying this scenario to explain the C/O$`>1`$ ratios observed in R-stars. The same figure can be applied in the case of J-stars. This extreme C/O ratio is not observed in any C-star. Furthermore, even assuming that the material transferred were Li-rich (some N-type carbon stars are Li-rich), it is unlikely that Li could survive during the mass-transfer and posterior mixing. In fact, extrinsic (binary, no Tc) S stars do not usually show the Li enhancements found here (Barbuy et al., 1992). Nevertheless, a significant number of J-stars ($`5\%10\%`$, LLoyd-Evans 1991) show a very uniform 9.85 $`\mu `$m emission which is believed to be due to the presence of a silicate dust shell (Little-Marenin, 1986). Also the detection of H<sub>2</sub>O masers in five J-type stars has been reported (Engels, 1994). This is rather strange, because silicate emission and H<sub>2</sub>O masers are usually associated with O-rich environments, while J-stars are C-rich objects. It is difficult to interpret these observations. The circumstellar shell is assumed to be produced by mass loss from the central carbon star and should reflect the carbon-rich material of its photosphere. The transit time of material through the circumstellar shell is only of the order of years, which is shorter than the transit time ($`10^5`$ yr) from M$``$S$``$C. Thus, it is unlikely that we are observing very recently produced circumstellar material from a progenitor M star. Our stars were classified as J-type C-stars over 50 years ago. It has been suggested (Little-Marenin, 1986; Lloyd-Evans, 1990, 1991) that the material expelled from the now carbon star, starting while it still had an oxygen-rich envelope, has accumulated in a disc (or common envelope) around an unseen hypothetical companion. Lambert, Hinkle & Smith (1990) discuss in detail the different possibilities for the nature and mass of the companion star, although the most probable situation is that the secondary is a low mass star on the main sequence. In fact, Barnbaum et al. (1991) found radial velocity variations by 6 km s<sup>-1</sup> in BM Gem and by 5 km s<sup>-1</sup> in EU And over a 6 month interval. The small uncertainty in their radial velocity measurements ($`\pm 1.5`$ km s<sup>-1</sup>) and the fact that the velocity has change of direction over this period of time, point out to a binary nature for these two J-type carbon stars. Although the binary hypothesis can probably explain the silicate emission in some J-stars, unfortunately there are not other radial velocity variation studies nor a search for ultraviolet excesses (in the hypothesis that the companion is now a white dwarf) to test the binary scenario for all the observed J-stars. Note that McClure (1997) found no evidence of binary motion in a sample of 22 R-type C-stars (possible ancestors of J-stars). He however, concluded that since it is very common to find binary systems ($`20\%`$) among normal late-type giants, it is likely that the R-stars were once all binaries, but with separations so close that would not allow them to evolve completely up the giant and asymptotic giant branches without coalescing. The star with the smallest mass in the system was disrupted and engulfed by the now visible J-star. Perhaps a strong non-standard mixing event between the core and the envelope material at the He-flash was induced by tidal forces in the actual J-star. Whether this stellar merging or non-standard He-flash triggered by binarity are able to induce the chemical properties observed in J-stars is an open question. ## 6 Concluding Remarks Our most important conclusion is that heavy element abundances in J-type carbon stars are nearly solar with respect to their metallicity. We did not found Tc in these stars, although we set some generous upper limits for two of the stars studied. Considering all our abundance results, it is difficult to find an evolutionary status for J-stars. Their average luminosity and variability types leads us to consider these objects as less evolved than normal (N) carbon stars. However, standard AGB models are unable to explain all their properties. On the contrary, the chemical peculiarities of J-stars suggests the existence of a non-standard mixing mechanism, similar to that proposed in the red giant branch to explain anomalous CNO isotopic ratios and Li abundances. This extra-mixing mechanism, acting preferably in the early AGB phase of low-mass stars (M$`23`$ M), would take material from the convective envelope, transport it down to regions hot enough for some nuclear processing and then transport it back up to the convective envelope. The expected stellar mass for the occurrence of this cool processing would be in agreement with the observational evidence suggesting a low-mass for most of the J-stars studied here. Mixing at the He-core flash and the binary system hypothesis may well be alternative scenarios, although fine tuning is required to explain all the observed characteristics of J-stars within these models. Nevertheless, these scenarios require further investigation. On the other hand, the existence of rather luminous J-stars (M$`{}_{\mathrm{bol}}{}^{}<5.5`$) in our galaxy, as well as in other galaxies (M31 and the Magellanic Clouds, see Brewer et al. 1996; Smith et al. 1995; Bessell, Wood and Lloyd-Evans 1983), suggests there could be two types of J-stars depending upon the initial mass of the parent star. The low-mass J-stars would be explained by a non-standard mixing mechanism such as those mentioned above, while the high initial mass (M$`4`$ M) J-stars (perhaps WZ Cas is an example of this) would be explained through the operation of HBB. This idea was previously proposed by Lorenz-Martins (1996). Note that current models of s-process nucleosynthesis (e.g. Busso & Gallino 1997) predict a strong metallicity dependence of s-nuclei enrichment. For solar metallicity and/or slightly metal-rich intermediate mass AGB stars, a small s-process element enrichment is predicted. This might be in agreement with the small \[$`<h>`$/M\] value that we have found in WZ Cas. The study of the presence of s-elements in the luminous J-stars of M31 and the Magellanic Clouds would be of a great interest. Data from the VALD data base at Vienna were used for the preparation of this paper. K. Eriksson and the stellar atmosphere group of the Uppsala Observatory are thanked for providing the grid atmospheres. The 4.2 m WHT is operated on the island of La Palma by the RGO in the Spanish Observatorio del Roque de los Muchachos of the Instituto de Astrofísica de Canarias. Based in part on observations collected at the German-Spanish Astronomical Centre, Carlar Alto, Spain. This work was partially supported by grant PB96-1428.
warning/0001/cond-mat0001126.html
ar5iv
text
# 2 Introduction ## 2 Introduction Ordering of complex systems has attracted interest of researchers working in the field of numerical simulations. Well-known examples of such complex systems may be a variety of glassy systems including window glasses, orientational glasses of molecular crystals, vortex glasses in superconductors and spin-glass magnets. Often, in the dynamics of such complex systems, characteristic slow relaxation is known to occur. It has been a great challenge for researchers in the field to clarify the nature and the origin of these slow dynamics, as well as to get fully equilibrium properties by overcoming the slow relaxation. In particular, spin glasses are the most extensively studied typical model system, for which numerous analytical, numerical and experimental works have been made.<sup>?</sup> Studies on spin glasses now have more than twenty years of history. Main focus of earlier studies was put on obtaining the equilibrium properties of spin glasses. While extensive experimental studies have clarified that the spin-glass magnets exhibit an equilibrium phase transition at a finite temperature,<sup>?</sup> the true nature of the experimentally observed spin-glass transition and that of the low-temperature spin-glass phase still remain open problems. It has been known that the magnetic interactions in many of real spin-glass materials are Heisenberg-like, in the sense that the magnetic anisotropy is much weaker than the exchange energy. In apparent contrast to experiments, numerical simulations have indicated that isotropic Heisenberg spin glass exhibits only a zero-temperature transition.<sup>?,?,?,?,?,?,?</sup> Apparently, there is a puzzle here: How can one reconcile the absence of the spin-glass order in an isotropic Heisenberg spin glass with the experimental observation? In order to solve this apparent puzzle, a chirality mechanism of experimentally observed spin-glass transitions was recently proposed by one of the authors,<sup>?,?</sup> on the assumption that an isotropic 3D Heisenberg spin glass exhibited a finite-temperature chiral-glass transition without the conventional spin-glass order, in which only spin-reflection symmetry was broken with preserving spin-rotation symmetry. “Chirality” is an Ising-like multispin variable representing the sense or the handedness of the noncoplanar spin structures. It was argued that, in real spin-glass magnets, the spin and the chirality were “mixed” due to the weak magnetic anisotropy and the chiral-glass transition was then “revealed” via anomaly in experimentally accessible quantities. Meanwhile, theoretical question whether there really occurs such finite-temperature chiral-glass transition in an isotropic 3D Heisenberg spin glass, a crucial assumption of the chirality mechanism, remains somewhat inconclusive.<sup>?,?</sup> More recently, there arose a growing interest both theoretically and experimentally in the off-equilibrium dynamical properties of spin glasses. In particular, aging phenomena observed in many spin glasses<sup>?</sup> have attracted attention of researchers.<sup>?,?</sup> Unlike systems in thermal equilibrium, relaxation of physical quantities depends not only on the observation time $`t`$ but also on the waiting time $`t_w`$, i.e., how long one waits at a given state before the measurements. Recent studies have revealed that the off-equilibrium dynamics in the spin-glass state generally has two characteristic time regimes.<sup>?,?,?</sup> One is a short-time regime, $`t_0tt_w`$ ($`t_0`$ is a microscopic time scale), called “quasi-equilibrium regime”, and the other is a long-time regime, $`tt_w`$, called “aging regime” or “out-of-equilibrium regime”. In the quasi-equilibrium regime, the relaxation is stationary and the fluctuation-dissipation theorem (FDT) holds. On the other hand, in the aging regime, the relaxation becomes non-stationary, FDT broken. Although theoretical studies of off-equilibrium dynamics on Ising-like models<sup>?,?,?,?</sup> succeeded in reproducing some of the features of experimental results, it is clearly desirable to study the dynamical properties of a Heisenberg spin-glass model to make a direct link between theory and experiments. In the present article, we report on our recent results of equilibrium as well as off-equilibrium Monte Carlo (MC) simulations on an isotropic 3D Heisenberg spin glass, with emphasis on their dynamical aspects. Ordering properties of both the spin and the chirality are studied, aimed at testing the validity of the proposed chirality scenario of spin-glass transitions. We note that MC simulation is particularly suited to this purpose, since, at the moment, chirality itself is not directly measurable experimentally. By contrast, it is straightforward to measure the chirality by numerical simulations. ## 3 Model The model we simulate is the classical Heisenberg model on a simple cubic lattice with the nearest-neighbor random Gaussian couplings, $`J_{ij}`$, defined by the Hamiltonian $$=\underset{<ij>}{}J_{ij}𝐒_i𝐒_j,$$ (3.1) where $`𝐒_i=(S_i^x,S_i^y,S_i^z)`$ is a three-component unit vector, and the sum runs over all nearest-neighbor pairs with $`N=L\times L\times L`$ spins. $`J_{ij}`$ is the isotropic exchange coupling with zero mean and variance $`J`$. Frustration in vector spin systems often causes noncollinear or noncoplanar spin structures. Such noncollinear or noncoplanar orderings give rise to a nontrivial chirality according as the spin structure is either right- or left-handed.<sup>?</sup> We define the local chirality at the $`i`$-th site and in the $`\mu `$-th direction, $`\chi _{i\mu }`$, for three Heisenberg spins by, $$\chi _{i\mu }=𝐒_{i+\widehat{𝐞}_\mu }(𝐒_i\times 𝐒_{i\widehat{𝐞}_\mu }),$$ (3.2) where $`\widehat{𝐞}_\mu (\mu =x,y,z)`$ denotes a unit lattice vector along the $`\mu `$-axis. The chirality is a pseudoscalar in the sense that it is invariant under global spin rotation but changes sign under global spin reflection. Chiral order is related to a possible breaking of the spin-reflection symmetry with preserving the spin-rotation symmetry. ## 4 Equilibrium simulations First, we report on our equilibrium MC simulations of a fully isotropic 3D Heisenberg spin glass defined by Eq. (3.1).<sup>?</sup> Monte Carlo simulation is performed based on an exchange method developed by Hukushima and Nemoto,<sup>?</sup> where whole configurations at two neighboring temperatures of the same sample are occasionally exchanged keeping the system at equilibrium. The method has turned out to be efficient in thermalizing the system showing slow relaxation over free energy barriers.<sup>?</sup> By this method, we succeed in equilibrating the system down to the temperature considerably lower than those attained in the previous simulations. We run in parallel two independent replicas with the same bond realization and compute an overlap between the chiral variables in the two replicas, $$q_\chi =\frac{1}{3N}\underset{i,\mu }{}\chi _{i\mu }^{(1)}\chi _{i\mu }^{(2)}.$$ (4.1) The lattice sizes studied are $`L=6,8,10,12,16`$ with periodic boundary conditions. In the case of $`L=16`$, for example, we prepare 60 temperature points distributed in the range \[$`0.10J,0.25J`$\] for a given sample, and perform $`1.2\times 10^6`$ exchanges per temperature of the whole lattices combined with the same number of standard single-spin-flip heat-bath sweeps. For $`L=16`$, we equilibrate the system down to the temperature $`T/J=0.10`$, which is lower than the minimum temperature attained previously. Sample average is taken over 1500 ($`L=6`$), 1200 ($`L=8`$), 640 ($`L=10`$), 296 ($`L=12`$) and 136 ($`L=16`$) independent bond realizations. Equilibration is checked by monitoring the stability of the results against at least three-times longer runs for a subset of samples. Information about the equilibrium dynamics can be obtained via the spin and chirality autocorrelation functions defined by $`C_s(t)`$ $`=`$ $`{\displaystyle \frac{1}{N}}{\displaystyle \underset{i}{}}[𝐒_i(t_0)𝐒_i(t+t_0)],`$ (4.2) $`C_\chi (t)`$ $`=`$ $`{\displaystyle \frac{1}{3N}}{\displaystyle \underset{i,\mu }{}}[\chi _{i\mu }(t_0)\chi _{i\mu }(t+t_0)],`$ (4.3) where $`\mathrm{}`$ and $`[\mathrm{}]`$ represent the thermal average and the average over bond disorder, respectively, while $`t_0`$ denotes some initial time at which the system is already fully equilibrated. Although we have used the exchange MC method for thermalizing the system, the time evolution during the measurements of autocorrelations has been made via the standard single-spin-flip heat-bath updating. MC time dependence of the calculated $`C_s(t)`$ and $`C_\chi (t)`$ are shown in Figs. 4 on log-log plots for several temperatures. As can be seen from Fig. 4(a), $`C_s(t)`$ shows a downward curvature at all temperature studied, suggesting an exponential-like decay characteristic of the disordered phase, consistently with the absence of the standard spin-glass order. In sharp contrast to this, $`C_\chi (t)`$ shows either a downward curvature characteristic of the disordered phase, or an upward curvature characteristic of the long-range ordered phase, depending on whether the temperature is higher or lower than $`T/J0.16`$, while just at $`T/J0.16`$ the linear behavior corresponding to the power-law decay is observed: See Fig. 4(b). Hence, our dynamical data indicates that the chiral-glass order without the standard spin-glass order takes place at $`T_{\mathrm{CG}}/J=0.160\pm 0.005`$, below which a finite chiral Edwards-Anderson (EA) order parameter $`q_{\mathrm{CG}}^{\mathrm{EA}}>0`$ develops. From the slope of the data at $`T=T_{\mathrm{CG}}`$, the exponent $`\lambda `$ characterizing the power-law decay of $`C_\chi (t)t^\lambda `$ is estimated to be $`\lambda =0.193\pm 0.005`$. We note that the occurrence of a chiral-glass transition was also supported by the behaviors of the static Binder ratios (not shown here).<sup>?</sup> Establishing the existence of a finite-temperature chiral-glass transition, we proceed to the study of the properties of the chiral-glass ordered state itself. In Fig. 4, we display the distribution function of the chiral-overlap defined by $$P(q_\chi ^{})=[\delta (q_\chi q_\chi ^{})],$$ (4.4) calculated by the exchange MC method at a temperature $`T/J=0.1`$, well below the chiral-glass transition temperature. As is evident from Fig. 4, the shape of the calculated $`P(q_\chi )`$ is somewhat different from the one observed in the standard Ising-like models such as the 3D EA model or the mean-field SK model. As usual, $`P(q_\chi )`$ has standard “side-peaks” corresponding to the EA order parameter $`\pm q_{\mathrm{CG}}^{\mathrm{EA}}`$, which grow and sharpen with increasing $`L`$. In addition to the side peaks, an unexpected “central peak” at $`q_\chi =0`$ shows up for larger $`L`$, which also grows and sharpens with increasing $`L`$. This latter aspect, i.e., the existence of a central peak growing and sharpening with the system size, is a peculiar feature of the chiral-glass ordered phase never observed in the EA or the SK models. Since we do not find any sign of a first-order transition such as a discontinuity in the energy, the specific-heat nor the order parameter $`q_{\mathrm{CG}}^{\mathrm{EA}}`$, this feature is likely to be related to a nontrivial structure in the phase space associated with the chirality. We note that this peculiar feature is reminiscent of the behavior characteristic of some mean-field models showing the so-called one-step replica-symmetry breaking (RSB).<sup>?</sup> In Figs. 4, we show the (exchange) MC time dependence of $`P(q_\chi )`$, together with those of the height of the central peak $`P(0)`$ and of the side peak $`P(q_{\mathrm{EA}})`$. These quantities are found to reach stationary values exponentially fast with the correlation time of order of $`10^5`$ MC steps, while actual spin configurations continue to change via the temperature-exchange process and the relatively rapid relaxation realized at higher temperatures. ## 5 Off-equilibrium simulations In this section, we report on our results of off-equilibrium MC simulations.<sup>?</sup> Unlike the case of equilibrium simulations, the system here is never in full thermal equilibrium. Recent studies have revealed that one can still get useful information from such off-equilibrium simulations, even including certain equilibrium properties. The quantities we are mainly interested here are the off-equilibrium spin and chirality autocorrelation functions defined by $`C_s(t_w,t+t_w)`$ $`=`$ $`{\displaystyle \frac{1}{N}}{\displaystyle \underset{i}{}}[𝐒_i(t_w)𝐒_i(t+t_w)],`$ (5.1) $`C_\chi (t_w,t+t_w)`$ $`=`$ $`{\displaystyle \frac{1}{3N}}{\displaystyle \underset{i,\mu }{}}[\chi _{i\mu }(t_w)\chi _{i\mu }(t+t_w)].`$ (5.2) MC simulation is performed based on the standard single spin-flip heat-bath method here. Starting from completely random initial configurations, the system is quenched to a working temperature. Total of about $`3\times 10^5`$ MC steps per spin are generated in each run. Sample average is taken over 30-120 independent bond realizations, four independent runs being made using different spin initial conditions and different sequences of random numbers for each sample. The lattice size mainly studied is $`L=16`$ with periodic boundary conditions, while in some cases lattices with $`L=12`$ and $`24`$ are also studied. The spin and chirality autocorrelation functions at a low temperature $`T/J=0.05`$ are shown in Fig. 5 as a function of $`t`$. For larger $`t_w`$, the curves of the spin autocorrelation function $`C_s`$ come on top of each other in the long-time regime, indicating that the stationary relaxation is recovered and aging is interrupted. This behavior has been expected because the 3D Heisenberg spin glass has no standard spin-glass order.<sup>?,?,?,?,?,?,?</sup> By contrast, the chiral autocorrelation function $`C_\chi `$ shows an entirely different behavior: Following the initial decay, it exhibits a clear plateau around $`tt_w`$ and then drops sharply for $`t>t_w`$. It also shows an eminent aging effect, namely, as one waits longer, the relaxation becomes slower and the plateau-like behavior at $`tt_w`$ becomes more pronounced. It should be noticed that the plateau-like behavior observed here has been hardly noticeable in simulations of the 3D EA model. While the plateau-like behavior observed in $`C_\chi `$ is already suggestive of a nonzero chiral Edwards-Anderson order parameter, $`q_{\mathrm{CG}}^{\mathrm{EA}}>0`$, more quantitative analysis similar to the one recently done by Parisi et al for the 4D Ising spin glass<sup>?</sup> is performed to extract $`q_{\mathrm{CG}}^{\mathrm{EA}}`$ from the data of $`C_\chi `$ in the quasi-equilibrium regime. Finiteness of $`q_{\mathrm{CG}}^{\mathrm{EA}}`$ is also visible in a log-log plot of $`C_\chi `$ versus $`t`$ as shown in the inset of Fig. 5, where the data show a clear upward curvature. We extract $`q_{\mathrm{CG}}^{\mathrm{EA}}`$ by fitting the data of $`C_\chi `$ for $`t_w=3\times 10^5`$ to the power-law form, $`C(t_w,t+t_w)=q^{\mathrm{EA}}+\frac{C}{t^\lambda }`$, in the time range $`40t3,000`$ satisfying $`t/t_w0.01`$. The obtained $`q_{\mathrm{CG}}^{\mathrm{EA}}`$, plotted as a function of temperature in Fig. 5, clearly indicates the occurrence of a finite-temperature chiral-glass transition at $`T_{\mathrm{CG}}/J=0.157\pm 0.01`$ with the associated order-parameter exponent $`\beta _{\mathrm{CG}}=1.1\pm 0.1`$. The size dependence turns out to be rather small, although the mean values of $`q_{\mathrm{CG}}^{\mathrm{EA}}`$ tend to slightly increase around $`T_{\mathrm{CG}}`$ with increasing $`L`$. Since both finite-size effect and finite-$`t_w`$ effect tend to underestimate $`q_{\mathrm{CG}}^{\mathrm{EA}}`$, one may regard the present result as a rather strong evidence of the occurrence of a finite-temperature chiral-glass transition. The present estimate of the transition temperature $`T_{\mathrm{CG}}/J=0.157\pm 0.01`$ is in very good agreement with the equilibrium estimate in the preceding section, $`T_{\mathrm{CG}}/J=0.160\pm 0.005`$. If we combine the present estimate of $`\beta _{\mathrm{CG}}`$ with the estimate of $`\gamma _{\mathrm{CG}}`$ in the preceding section and use the scaling relations, various chiral-glass exponents can be estimated to be $`\alpha 1.7`$, $`\beta _{\mathrm{CG}}1.1`$, $`\gamma _{\mathrm{CG}}1.5`$, $`\nu _{\mathrm{CG}}1.2`$ and $`\eta _{\mathrm{CG}}0.8`$. The dynamical exponent is estimated to be $`z_{\mathrm{CG}}4.7`$ by using the estimated value of $`\lambda `$ and the scaling relation $`\lambda =\beta _{\mathrm{CG}}/z_{\mathrm{CG}}\nu _{\mathrm{CG}}`$. While the dynamical exponent $`z_{\mathrm{CG}}`$ comes rather close to the $`z`$ of the 3D Ising EA model, The obtained static exponents differ significantly from those of the 3D Ising EA model $`\beta 0.6`$, $`\gamma 4`$, $`\nu 2`$ and $`\eta 0.35`$,<sup>?,?,?,?,?</sup> suggesting that the universality class of the chiral-glass transition of the 3D Heisenberg spin glass differs from that of the standard 3D Ising spin glass. According to the chirality mechanism, the criticality of real spin-glass transitions is derived from that of the chiral-glass transition of an isotropic Heisenberg spin glass, so long as the magnitude of random anisotropy is not too strong. If one tentatively accepts this scenario, the present result opens up an interesting possibility that the universality class of many of real spin-glass transitions might differ from that of the standard Ising spin glass, contrary to common belief. ## 6 Conclusion In summary, spin-glass and chiral-glass orderings of isotropic 3D Heisenberg spin glasses are studied by extensive MC simulations. Clear evidence of the occurrence of a finite-temperature chiral-glass transition without the conventional spin-glass order is presented both by equilibrium and off-equilibrium simulations. Spin and chirality show very different dynamical behaviors consistent with the “spin-chirality separation”. While the spin autocorrelation exhibits only an interrupted aging, the chirality autocorrelation persists to exhibit a pronounced aging effect reminiscent of the one observed in the mean-field model. The universality class of the chiral-glass transition is different from that of the the standard Ising spin glass, while the chiral-glass ordered state appears to exhibit a feature of “one-step” replica-symmetry breaking. We expect that these numerical findings have important implications to the understanding of the nature of real spin-glass ordering. Acknowledgment We would like to thank Prof. I. Campbell, Prof. A. P. Young and Prof. H. Takayama for useful discussion. The numerical calculation was performed on the Fujitsu VPP500 at the supercomputer center, ISSP, University of Tokyo, on the Hitachi SR2201 at the supercomputer center, University of Tokyo, and on the CP-PACS computer at the Center for Computational Physics, University of Tsukuba. References
warning/0001/cond-mat0001341.html
ar5iv
text
# Laser probing of atomic Cooper pairs ## Abstract We consider a gas of attractively interacting cold Fermionic atoms which are manipulated by laser light. The laser induces a transition from an internal state with large negative scattering length to one with almost no interactions. The process can be viewed as a tunneling of atomic population between the superconducting and the normal states of the gas. It can be used to detect the BCS-ground state and to measure the superconducting order parameter. Successful cooling of trapped Fermionic <sup>40</sup>K atoms down to temperatures where the Fermi degeneracy sets in was reported recently . This breakthrough opens up new opportunities for studying fundamental quantum statistical and many-body physics. A major advantage of Fermionic atoms compared to electrons in condensed matter is the richness of their internal energy structure and the possibility to accurately and efficiently manipulate these energy states by laser light and magnetic fields. Furthermore, atomic gases are dilute and weakly interacting thus offering the ideal tool for developing and experimentally testing theories of many-body quantum physics. A major goal is to observe the predicted BSC-transition for Fermionic atoms – this would compare to the experimental realization of atomic Bose-Einstein condensates . It is still, however, an open question how to observe the BCS-transition, because the value of the superconducting order parameter (gap) is expected to be small and electro-magnetic phenomena such as the Meissner effect do not take place. We propose a method to detect the existence of the BCS-ground state and to measure the gap using laser light. The main difference in our method compared to the proposals of using off-resonant light scattering is that the light is resonant, that is, population is transferred from one internal state to another. Furthermore, the final state of this process is chosen to have negligible scattering length compared to the initial one: this effectively creates a superconducting – normal state interface across which the atomic population can move. There is a conceptual analogy to electron tunneling from a superconducting metal to a normal one, although the physical systems and their describtions differ in an essential way. Also non-optical phenomena, such as collective and single particle excitations, have been proposed to be used for observing the BCS-transition . The specific feature of our method with respect to other suggested probes, both optical and mechanical (modulating the trap frequency), is the utilization of the superconducting – normal state interface. The advantage is that the normal state population can be initally zero, thus even a small change in it is a significant relative effect. Furthermore, our work offers a basic starting point for describing any phenomenon related to the superconducting–normal interface. For instance, our method can be understood as an outcoupler. With small modifications, one could also describe outcoupling of pairs instead of single atoms which would create an atomic beam with BCS-type statistics. We consider atoms with three internal states available, say $`|e`$, $`|g`$, and $`|g^{}`$. They are chosen so that the interaction between atoms in states $`|g`$ and $`|g^{}`$ is relatively strong and attractive, and all other interactions are negligible. The laser frequency is chosen to transfer population between $`|e`$ and $`|g`$, but is not in resonance with any transition which could move population away from the state $`|g^{}`$. If there is more or less equal amount of atoms in states $`|g`$ and $`|g^{}`$ they can form the superconducting BCS ground state; atoms in $`|e`$ are in the normal state. For small intensities, the laser-interaction can be treated as a perturbation, the unperturbed states being the normal and the superconducting state. The transfer of atoms from $`|g`$ to $`|e`$ is then analogous to tunneling of electrons from a normal metal to a superconductor induced by an external voltage, which can be used as a method to probe the gap and the density of states of the superconductor . In our case the tunneling is between two internal states rather than two spatial regions, this resembles the idea of internal Josephson-oscillations in two-component Bose-Einstein condensates . Figure 1 illustrates the basic idea. The observable carrying essential information about the superconducting state is the change in the population of the state $`|e`$, we call this the current $`I`$. Below we calculate the current first in the homogeneous case and then assuming that the atoms are confined in a harmonic trap. Finally will discuss possibilities for experimental realization of the idea, and choosing the states $`|e`$, $`|g`$, and $`|g^{}`$ for <sup>6</sup>Li. We define a two-component Fermion field $`\psi (𝐱)=(\psi _e(𝐱)\psi _g(𝐱))^T`$, where $`\psi _e`$ and $`\psi _g`$ fulfil standard Fermionic commutation relations. The fields $`\psi _{e/g}`$ can be expanded using some basis functions (e.g. plane waves or trap wave functions) and corresponding creation and annihilation operators: $`\psi _{e/g}(𝐱)=_jc_j^{e/g}\varphi _j^{e/g}(𝐱)`$. The annihilation and creation operators fulfil $`\{c_{i}^{e}{}_{}{}^{},c_j^g\}=0`$ and $`\{c_i^{e/g},c_j^{e/g}\}=\delta _{ij}`$. The two components of the field, corresponding to the internal states $`|e`$ and $`|g`$, are coupled by a laser. This can be either a direct excitation or a Raman process; we denote the atomic energy level difference by $`\omega _a`$ ($`\mathrm{}1`$), the laser frequency $`\omega _L`$ and the wave vector $`k_L`$ – in the case of a Raman process these are effective quantities. In the rotating wave approximation the Hamiltonian reads $`H`$ $`=`$ $`H_e+H_{gg^{}}+{\displaystyle \frac{\mathrm{\Delta }}{2}}{\displaystyle d^3x\psi ^{}(𝐱)\sigma _z\psi (𝐱)}`$ (4) $`+{\displaystyle d^3x\psi ^{}(𝐱)\left(\begin{array}{cc}0& \mathrm{\Omega }(𝐱)\\ \mathrm{\Omega }^{}(𝐱)& 0\end{array}\right)\psi (𝐱)}.`$ Here $`\mathrm{\Delta }=\omega _a\omega _L`$ is the (effective) detuning, and $`\mathrm{\Omega }(𝐱)`$ contains the spatial dependence of the laser field multiplied with the (effective) Rabi frequency. The parts $`H_e`$ and $`H_{gg^{}}`$ contain terms which depend only on $`\psi _e`$ or $`\psi _g`$, $`\psi _g^{}`$, respectively. Possible spatial inhomogeneity, e.g. from the trap potential, is also included in $`H_e`$ and $`H_{gg^{}}`$. The observable carrying essential information about the superconducting state is the rate of change in population of the $`|e`$ state. We may call it, after the electron tunneling analogy, the current $`I(t)=\dot{N_e}`$, where $`N_e=d^3x\psi _e^{}(𝐱)\psi _e(𝐱)`$. The current $`I(t)`$ is calculated considering the tunneling part of the Hamiltonian, $`H_T=H(H_e+H_{gg^{}}+\mathrm{\Delta }/2(N_eN_g))`$ as a perturbation; the current $`I`$ becomes the first order response to the external perturbation caused by the laser. We calculate it both in the homogeneous case and in the case of harmonic confinement. The calculations are done in the grand canonical ensemble, therefore the chemical potentials $`\mu _g`$ and $`\mu _e`$ are introduced. Also the detuning $`\mathrm{\Delta }`$ acts like a difference in chemical potentials, thus it becomes useful to define an effective quantity of the form $`\stackrel{~}{\mathrm{\Delta }}=\mu _e\mu _g+\mathrm{\Delta }\mathrm{\Delta }\mu +\mathrm{\Delta }`$. In the derivation we assume finite temperature, but here we only quote the results for $`T=0`$. Homogeneous case: The assumption of spatial homogeneity is appropriate when the atoms are confined in a trap potential which changes very little compared to characteristic quantities of the system, such as the coherence length and the size of the Cooper pairs. In the homogeneous case we expand the Fermion fields $`\psi _{e/g}`$ into plane waves. The Hamiltonian becomes $`H`$ $`=`$ $`H_e+H_{gg^{}}+{\displaystyle \frac{\mathrm{\Delta }}{2}}{\displaystyle \underset{k}{}}[c_{k}^{e}{}_{}{}^{}c_k^ec_{k}^{g}{}_{}{}^{}c_k^g]`$ (6) $`+{\displaystyle \underset{kl}{}}[T_{kl}c_{k}^{e}{}_{}{}^{}c_l^g+h.c.],`$ where $`T_{kl}=\frac{1}{V}d^3x\mathrm{\Omega }(𝐱)e^{i𝐤𝐱}e^{i𝐥𝐱}`$. We calculate the current $`I=\dot{N_e}=i[H,N_e]`$ treating $`H_T`$ as a perturbation: terms of higher order than $`H_T^2`$ are neglected. Because we are interested in the current between the superconducting and normal states, correlations of the form $`c_e^{}c_e^{}c_gc_g`$ (and $`h.c.`$) are omitted since they correspond to tunneling of pairs (Josephson current). The current can be written $`I={\displaystyle _{\mathrm{}}^{\mathrm{}}}𝑑t\theta (t)(e^{i\stackrel{~}{\mathrm{\Delta }}t}[A^{}(0),A(t)]e^{i\stackrel{~}{\mathrm{\Delta }}t}[A(0),A^{}(t)])`$ (7) where $`A(t)=_{kl}T_{kl}c_{k}^{g}{}_{}{}^{}(t)c_l^e(t)`$, and $`c_l^{e/g}(t)=e^{iKt}c_l^{e/g}e^{iKt}`$ where $`K=H\mu _eN_e\mu _gN_g`$. The two terms in the above equation have the form of retarded and advanced Green’s functions. These are evaluated using Matsubara Green’s functions techniques, which leads to $`I=_{kl}|T_{kl}|^2_{\mathrm{}}^{\mathrm{}}\frac{dϵ}{2\pi }[n_F(ϵ)n_F(ϵ+\stackrel{~}{\mathrm{\Delta }})]A_g(k,ϵ+\stackrel{~}{\mathrm{\Delta }})A_e(l,ϵ),`$ where $`n_F`$ are the Fermi distribution functions and $`A_{g/e}`$ are the spectral functions for the superconducting and normal states. We use the standard expressions $`A_e(l,ϵ)=2\pi \delta (ϵ\xi _l)`$ and $`A_g(k,ϵ+\stackrel{~}{\mathrm{\Delta }})=2\pi [u_k^2\delta (ϵ+\stackrel{~}{\mathrm{\Delta }}\omega _k)+v_k^2\delta (ϵ+\stackrel{~}{\mathrm{\Delta }}+\omega _k)]`$ . Here $`\xi _l=E_l\mu _e`$ and $`u_k`$, $`v_k`$ and $`\omega _k`$ are given by the Bogoliubov transformation. Because we are interested in the tunneling out of the superconductor, only the term proportional to $`v_k^2`$ is considered. The laser field is chosen to be a running wave, that is $`\mathrm{\Omega }(𝐱)=\mathrm{\Omega }e^{i𝐤_𝐋𝐱}`$. The term $`|T_{kl}|^2`$ now produces a delta-function enforcing momentum conservation. Note that this is very different from the assumption of a constant transfer matrix ($`_{kl}|T_{kl}|^2|T|^2_{kl}`$) made in the standard calculation for tunneling of electrons over a superconductor – normal metal surface . The final result becomes (assuming $`T=0`$) $`I=2\pi \mathrm{\Omega }^2\rho \theta (\stackrel{~}{\mathrm{\Delta }}\omega _{\stackrel{~}{k}k_L}\mathrm{\Delta }\mu ){\displaystyle \frac{\omega _{\stackrel{~}{k}k_L}\xi _{\stackrel{~}{k}k_L}}{\omega _{\stackrel{~}{k}k_L}+\xi _{\stackrel{~}{k}k_L}\left[1\frac{k_L}{\stackrel{~}{k}}\right]}}`$ (8) where $`\stackrel{~}{k}`$ is given by the following energy conservation condition: $`\stackrel{~}{\mathrm{\Delta }}+\omega _{\stackrel{~}{k}k_L}+\xi _{\stackrel{~}{k}k_L}=0`$, $`\omega _k=\sqrt{\xi _k^2+\mathrm{\Delta }_G^2}`$, and $`\rho `$ is the density of states which appears when the summation over momenta is changed into an integration over energies. The laser momentum $`k_L`$ can be very small compared to the momentum of the atoms, especially in the case of a Raman process. By setting $`k_L=0`$ the result becomes (choosing $`\stackrel{~}{\mathrm{\Delta }}<0`$ i.e. current from $`|g`$ to $`|e`$) $`I=2\pi \mathrm{\Omega }^2\rho \theta (\stackrel{~}{\mathrm{\Delta }}^2\mathrm{\Delta }_G^2+2\mathrm{\Delta }\mu \stackrel{~}{\mathrm{\Delta }}){\displaystyle \frac{\mathrm{\Delta }_G^2}{\stackrel{~}{\mathrm{\Delta }}^2}}.`$ (9) To understand the results in terms of physics, let us first consider the case of equal chemical potentials $`\mathrm{\Delta }\mu =0`$: $`I=2\pi \mathrm{\Omega }^2\rho \theta (\mathrm{\Delta }\mathrm{\Delta }_G){\displaystyle \frac{\mathrm{\Delta }_G^2}{\mathrm{\Delta }^2}}.`$ (10) In order to transfer one atom from the state $`|g`$ to $`|e`$ the laser has to break a Cooper pair. The minimum energy required for this is the gap energy $`\mathrm{\Delta }_G`$, therefore the current does not flow before the laser detuning provides this energy – this is expressed by the step function in (10). As $`|\mathrm{\Delta }|`$ increases further, the current will decrease quadratically. This is because the case $`|\mathrm{\Delta }|=\mathrm{\Delta }_G`$ corresponds to the transfer of particles with $`p=p_F`$, whereas larger $`|\mathrm{\Delta }|`$ means larger momenta, and there are simply fewer Cooper-pairs away from the Fermi surface. This behaviour is very different from the electron tunneling where the current grows as $`\sqrt{(eV)^2\mathrm{\Delta }_G^2}`$ (the voltage $`eV`$ corresponds to the detuning $`\mathrm{\Delta }`$ in our case) because all momentum states are coupled to each other. When the chemical potentials are not equal the situation is more complicated, but the basic features are the same: i) treshold for the onset of the current given basically by the gap energy, and ii) further decay of the current because the density of the states that can fulfil energy and momentum conservation decreases. Harmonic confinement: The trap wave functions provide the natural basis of expansion when the atoms are confined in a harmonic trap. We will not, however, do this expansion from the beginning of the calculations but rather derive the current and the corresponding (position dependent) Green’s functions directly for the Fermion fields $`\psi _{e/g}`$. The current becomes $`I=2Im[X_{ret}(\stackrel{~}{\mathrm{\Delta }})]`$, $`X_{ret}(\stackrel{~}{\mathrm{\Delta }})={\displaystyle _{\mathrm{}}^{\mathrm{}}}{\displaystyle \frac{dϵ}{2\pi }}{\displaystyle d^3xd^3x^{}\mathrm{\Omega }^{}(𝐱)\mathrm{\Omega }(𝐱^{})}`$ (11) $`[\stackrel{~}{A}_e(𝐱,𝐱^{},ϵ)G_{adv}^g(𝐱^{},𝐱,ϵ+\stackrel{~}{\mathrm{\Delta }})`$ (12) $`+G_{ret}^e(𝐱,𝐱^{},ϵ\stackrel{~}{\mathrm{\Delta }})\stackrel{~}{A}_g(𝐱^{},𝐱,ϵ)],`$ (13) and $`\stackrel{~}{A}_{e/g}`$ are defined $`\stackrel{~}{A}_{e/g}(𝐱,𝐱^{},ϵ)=i(G_{ret}^{e/g}(𝐱,𝐱^{},ϵ)G_{adv}^{e/g}(𝐱,𝐱^{},ϵ))`$. In order to take the imaginary part of the expression (13) we use the following Green’s functions $`G_{adv}^g(𝐱^{},𝐱,ϵ)`$ $`=`$ $`{\displaystyle \underset{n}{}}{\displaystyle \frac{u_n(𝐱^{})u_n^{}(𝐱)}{ϵ\omega _ni\delta }}+{\displaystyle \frac{v_n^{}(𝐱^{})v_n(𝐱)}{ϵ+\omega _ni\delta }}`$ (14) $`G_{ret}^e(𝐱,𝐱^{},ϵ)`$ $`=`$ $`{\displaystyle \underset{n}{}}{\displaystyle \frac{\varphi _n(𝐱)\varphi _n^{}(𝐱^{})}{ϵ\xi _n+i\delta }}.`$ (15) Here $`u_n(𝐱)`$, $`v_n(𝐱)`$ and $`\omega _n`$ are given by the Bogoliubov–DeGennes equations . We also use the fact that the trap wave functions $`\varphi _m(𝐱)`$ are real. The derivation gives $`I`$ $`=`$ $`2\pi {\displaystyle \underset{n,m}{}}\left|{\displaystyle d^3x\mathrm{\Omega }(𝐱)u_n(𝐱)\varphi _m(𝐱)}\right|^2[n_F(\omega _n)n_F(\xi _m)]`$ (18) $`\delta (\xi _m+\stackrel{~}{\mathrm{\Delta }}\omega _n)+\left|{\displaystyle d^3x\mathrm{\Omega }(𝐱)v_n^{}(𝐱)\varphi _m(𝐱)}\right|^2`$ $`[n_F(\omega _n)n_F(\xi _m)]\delta (\xi _m+\stackrel{~}{\mathrm{\Delta }}+\omega _n).`$ Again, we consider only the term proportional to $`v_n`$, assume zero temperature and use the known form for the trap energy $`\xi _m=m\mathrm{\Omega }_t\mu _e`$ to obtain the final result $`I={\displaystyle \frac{2\pi }{\mathrm{\Omega }_t}}{\displaystyle \underset{n}{}}\theta (\stackrel{~}{\mathrm{\Delta }}+\omega _n)\left|{\displaystyle d^3x\mathrm{\Omega }(𝐱)v_n^{}(𝐱)\varphi _M(𝐱)}\right|^2,`$ (19) where the quantum number $`M=(\mu _e\stackrel{~}{\mathrm{\Delta }}+\omega _n)/\mathrm{\Omega }_t`$. Although the result is not as transparent as in the homogeneous case, the characteristic treshold behaviour is expressed by the step function. Since the functions inside the absolute square are typically oscillating ones, this term would restrict the values of $`n`$ in the summation. We have checked that with strong approximations one obtains again the homogeneous case result. Experimental realization: Typical experimental parameters in the case of <sup>6</sup>Li are $`\mathrm{\Omega }_t/2\pi 150Hz`$ for the trap frequency and $`\mu _g100kHz`$ for the chemical potential . The order of magnitude of the gap is $`\mathrm{\Delta }_G/ϵ_F0.1`$ (see e.g. ). In Figure 2 we show the dependence of the current on the detuning in the homogeneous case, for $`\mathrm{\Delta }_G=0.1\mu _g`$. The sharp peak given by the result of Eq.(9) will be broadened in an experimental situation. One of the main causes of broadening is that it is difficult to fix the particle number with high accuracy. We have simulated this by assuming that the particle number ($`\mu _g`$) varies from experiment to experiment according to a Gaussian distribution (the width of the Gaussian is used as a measure of the fluctuations), and averaged over the different results. In these calculations we have introduced corresponding Gaussian fluctuations in the gap energy, because it is determined by the particle number. For as large as 10% fluctuations in the gap the peak is still clearly visible although broadened and slightly shifted. It is possible to prepare the desired initial number of atoms in states $`|e`$, $`|g`$ and $`|g^{}`$ very accurately by applying suitable incoherent laser pulses — this is one of the advantages of atomic Fermi gases compared to condensed matter systems. The number of atoms in the states $`|g`$ and $`|g^{}`$ should be nearly equal to produce the superconducting state, whereas the number of atoms in state $`|e`$ is not critical. In the simplest case the state $`|e`$ would be empty in the beginning. The use of our method then requires to know $`\mu _g`$ rather accurately, since it appears in the tunable argument $`\stackrel{~}{\mathrm{\Delta }}=\mathrm{\Delta }\mu +\mathrm{\Delta }=\mu _g+\mathrm{\Delta }`$ of the current $`I(\stackrel{~}{\mathrm{\Delta }})`$. If this is difficult, it might be better to use an incoherent pulse to prepare initially $`\mathrm{\Delta }\mu =\mu _e\mu _g=0`$. In this case the chemical potential dependence disappears from the expressions in the homogeneous case. Finally, we discuss how to choose the states $`|e`$, $`|g`$ and $`|g^{}`$ for a real atom (for instance <sup>6</sup>Li). It will probably be difficult to find states in which the atoms do not interact at all. A more likely solution is to choose the states so that the interaction between all of them is weak and repulsive, except between $`|g`$ and $`|g^{}`$ strong and attractive. Even a weak and attractive interaction instead of weak and repulsive would do, because the gap, critical temperature, and other essential BCS quantities have an exponential dependence on the scattering length, so the effect becomes easily negligible. Particularly <sup>6</sup>Li seems to offer a variety of interaction strengths between different states. As shown in , in a magnetic field of about $`B0.02T`$ the scattering lenght of several low field seeking hyperfine states is as strong as $`a_S2000a_0`$ whereas the scattering length between two high field seeking states is around zero. This is an example of the potential of modifying the scattering lengths by magnetic fields. In this case the trap should be optical in order to confine also the high field seekers. Alternatively, in a magnetic trap atoms in these states (corresponds to the non-interacting state $`|e`$ in our calculation) would simply fly out — this could even make the detection of the current easier. The detection of the atoms in the state $`|e`$, that is, the measurement of our basic observable $`I`$, could be done for instance by direct resonance fluorescence, or when a Raman process is used, indirectly from the intensity difference of the two Raman beams. In conclusion, we have proposed a method to observe the superconducting order parameter (gap) in a gas of cold Fermionic atoms. The idea is based on creating a normal phase – superconductor interface by coupling internal states of different scattering lengths by a laser. The advantage of our scheme is the sensitivity to even small currents of atomic population caused by the laser interaction. Furthermore, our results extend beyond measuring the order parameter: they can be used as a starting point in describing various processes related to the transfer of atoms between two gas components (in superconducting or normal states). If both of the internal states used were strongly interacting, one could perhaps observe Josepson tunneling of pairs from one superconductor to another. With small modifications, our results also describe the functioning of an outcoupler from a superconducting gas of Fermionic atoms. Acknowledgements We thank H.T.C. Stoof for useful discussions. PT acknowledges the support by the TMR programme of the European Commission (ERBFMBICT983061) and the Academy of Finland (projects 42588 and 47140).
warning/0001/cond-mat0001245.html
ar5iv
text
# Phase separation in systems with charge ordering ## I Introduction The problem of charge ordering (CO) in magnetic oxides attracts an attention of theorists since the discovery of Verwey transition in magnetite in the end of thirties . An early theoretical description of this phenomenon was given e.g. in . Recently this problem was reexamined in a number of papers in connection with the colossal magnetoresistance in manganites, see e.g. . The mechanisms stabilizing the CO state may be different: the Coulomb repulsion of charge carriers (the energy minimization requires keeping the carriers as far away as possible, similar to the Wigner crystallization), or the electron-lattice interaction leading to the effective repulsion of electrons at the nearest neighbor sites. In all cases, charge ordering can arise in systems with the mixed valence if the electron bandwidth is sufficiently small — large electron kinetic energy stabilizes the homogeneous metallic state. In real materials, in contrast to the Wigner crystallization, the underlying lattice periodicity determines the preferential types of CO. Thus, in the simplest bipartite lattice, to which belong the colossal magnetoresistance manganites of the type R<sub>1-x</sub>A<sub>x</sub>MnO<sub>3</sub> (R = La, Pr; A = Ca, Sr), or layered manganites R<sub>2-x</sub>A<sub>x</sub>MnO<sub>4</sub>, R<sub>2-2x</sub>A<sub>1+2x</sub>Mn<sub>2</sub>O<sub>7</sub>, the optimum conditions for the formation of the CO state exist for doping $`x=1/2`$. At such value of $`x`$ the concentrations of Mn<sup>3+</sup> and Mn<sup>4+</sup> are equal, and the simple checkerboard arrangement is possible. The most remarkable experimental fact here is that even at $`x1/2`$ (in the underdoped manganites, $`x<1/2`$) only the simplest version of charge ordering is experimentally observed with alternating checkerboard structure of occupied and empty sites in the basal plane . In other words, this structure corresponds to the doubling of the unit cell, whereas the more complicated structures with longer period (or even incommensurate structures) do not actually appear in this case. Then, the natural question arises: how could we redistribute the extra or missing electrons in the case of arbitrary doping level, keeping the superstructure the same as for $`x=1/2`$ ? To answer this question, the experimentalists introduced the concept of incipient charge ordered state corresponding to the distortion of long-range CO by microscopic metallic clusters . In fact, the existence of such a state implies a kind of phase separation. Note that the phase separation scenario in manganites is very popular now . Nowadays, there is a growing evidence suggesting that an interplay between the charge ordering and the tendency toward phase separation plays an essential role in the physics of materials with colossal magnetoresistance. In this paper, we consider a simple model, which allows us to clarify the situation at arbitrary doping. We include in this model both the Coulomb repulsion of electrons on the neighboring sites and the magnetic interactions responsible for the magnetic ordering of manganites. After demonstrating the instability of the system toward phase separation in certain ranges of doping, we also consider the simplest form of the phase separation — the formation of metallic droplets in an insulating matrix, estimate parameters of such droplets, and construct the phase diagram illustrating the interplay between charge ordering, magnetic ordering, and phase separation. One has to note that the mechanism of the charge ordering considered below (the Coulomb repulsion) is not the only one. The electron-lattice interaction can also play an important role, see e.g. . In application to manganites, one has to take into account also orbital and magnetic interactions . These may be important, in particular, to explain the fact that the charge ordering in half-doped perovskite manganites is a checkerboard one only in the basal plane, but it is “in-phase” in the $`c`$\- direction. However, the nature of such charge ordering is not clear yet and presents a separate problem: it is not evident that the magnetic interactions responsible for this stacking of $`\mathrm{𝑎𝑏}`$-planes in is indeed the dominant mechanism. Let us also emphasize that the charge ordering in manganites is often observed at higher temperatures than the magnetic ordering, and one has to look for a model not relying heavily on magnetic interactions. Note that in contrast to magnetic interactions, the Coulomb interaction is one of the important factors always present in the systems under consideration. Moreover, it has a universal nature and does not depend critically on specific features of a particular system. Consequently, our treatment can be applied also to other systems with the charge ordering such as magnetite Fe<sub>3</sub>O<sub>4</sub> , cobaltites , nickelates , etc. ## II The simplest model for charge ordering Let us consider a simple lattice model for charge ordering: $$\widehat{H}=t\underset{<i,j>}{}c_i^+c_j+V\underset{<i,j>}{}n_in_j\mu \underset{i}{}n_i,$$ (1) where $`t`$ is the hopping integral, $`V`$ is the nearest neighbor Coulomb interaction (similar $`nn`$ repulsion can be also obtained via the interaction with the breathing-type optical phonons); $`\mu `$ is the chemical potential, and $`c_i^+`$ and $`c_j`$ are one-electron creation and annihilation operators, $`n_i=c_i^+c_i`$. Symbol $`<i,j>`$ denotes the summation over the nearest-neighbor sites. Here, we omit for simplicity spin indices. We also assume the absence of a double occupancy in this model due to the strong onsite repulsion between electrons. The models of the type (1) with the $`nn`$ repulsion being responsible for the charge ordering are the most popular ones to describe this phenomenon, see e.g. and references therein. Hamiltonian (1) captures the main physical effects; if necessary, one can add to it some extra terms, which we will also do in Section V below. In the main part of our paper, we will always speak about electrons. However in application to real manganites we will mostly have in mind less than half-doped (underdoped) systems of the type R<sub>1-x</sub>A<sub>x</sub>MnO<sub>3</sub> with $`x<1/2`$. Thus, for a real system one has to substitute holes for our electrons. All the theoretical treatment definitely remains the same (from the very beginning we could define operators $`c`$ and $`c^+`$ in (1) as the operators of holes); we hope that it will not lead to any misunderstanding. We consider below the simplest case of square (2D) or cubic (3D) lattices, where for $`x=1/2`$ the simple two-sublattice ordering would take place. As mentioned in the Introduction, this is the case in layered manganites,whereas the ordering in 3D perovskite manganites is like this only in the basal plane, the ordering being “in-phase” in the c direction. To account for this behavior, apparently a more complicated model would be necessary. For the case $`n=1/2`$, the model (1) was analyzed in many papers; we follow the treatment of the Ref. . As mentioned above, the Coulomb repulsion (second term in (1)) stabilizes charge ordering in the form of checkerboard arrangement of occupied and empty sites, whereas the first term (band energy) opposes this tendency. At arbitrary values of electron density $`n`$, we shall at first consider a homogeneous CO solution and use the same ansatz as in , namely $$n_i=n[1+(1)^i\tau ].$$ (2) Such an expression implies the doubling of lattice periodicity, with the local densities $`n_1=n(1+\tau )`$ and $`n_2=n(1\tau )`$ at the neighboring sites. Note that at $`n=1/2`$ for a general form of electron dispersion without nesting, the CO state exists only at sufficiently strong repulsion $`V>2t`$ . The order parameter is $`\tau <1`$ for finite $`V/2t`$, and the ordering in general is not complete, i.e. an average electron density $`n_i`$ differs from zero or one even at $`T=0`$. We use the same coupled Green function approach as in , which yields $$\{\begin{array}{ccc}(E+\mu )G_1t_kG_2zVn(1\tau )G_1=\frac{1}{2\pi }& & \\ & & \\ (E+\mu )G_2t_kG_1zVn(1+\tau )G_2=0& & \end{array}$$ (3) where $`G_1`$ and $`G_2`$ are the Fourier transforms of the normal lattice Green functions $`G_{il}=c_ic_l^+`$ for sites $`i`$ and $`l`$ belonging respectively to one or different sublattices, $`z`$ is the number of nearest neighbors and $`t_k`$ is the Fourier transform of a hopping matrix element. While deriving (3), we performed a mean-field decoupling and replaced the averages $`c_i^+c_i`$ by the onsite densities $`n_i=n[1+(1)^i\tau ]`$. Solution of (3) leads to the following spectrum: $$E+\mu =Vnz\pm \sqrt{(Vn\tau z)^2+t_k^2}=Vnz\pm \omega _k.$$ (4) The spectrum defined by (4) resembles the spectrum of superconductor and hence the first term under the square root is analogous to the superconducting gap squared. In other words, we can introduce the charge-ordering gap by the formula $$\mathrm{\Delta }=Vn\tau z.$$ It depends upon density not only explicitly, but also via the density dependence of $`\tau `$. Hence, we get $$\omega _k=\sqrt{\mathrm{\Delta }^2+t_k^2}.$$ (5) Note that there is one substantial difference between the spectrum of charge ordered state (5) and superconducting state, namely here for $`n1/2`$ the chemical potential does not enter under the square root in (5) in contrast to the spectrum of superconductor where $$\omega _k=\sqrt{(t_k\mu )^2+\mathrm{\Delta }^2}.$$ Then, we can find the following expressions for the Green functions : $$\{\begin{array}{ccc}& & G_1=\frac{A_k}{E+\mu Vnz\omega _k+i0}+\frac{B_k}{E+\mu Vnz+\omega _k+i0}\\ & & \\ & & G_2=\frac{t_k}{2\omega _k}\frac{1}{2\pi }\left[\frac{1}{E+\mu Vnz\omega _k+i0}\frac{1}{E+\mu Vnz+\omega _k+i0}\right],\end{array}$$ (6) where $$A_k=\frac{1}{4\pi }\left(1\frac{\mathrm{\Delta }}{\omega _k}\right),\text{ }\text{ }B_k=\frac{1}{4\pi }\left(1+\frac{\mathrm{\Delta }}{\omega _k}\right).$$ (7) After the standard Wick transformation $`E+i0iE`$ in the expression for $`G_1`$, we can find the densities in the following form $`n_1=n(1+\tau )={\displaystyle \left[\left(1\frac{\mathrm{\Delta }}{\omega _k}\right)f_F(\omega _k\mu +Vnz)+\left(1+\frac{\mathrm{\Delta }}{\omega _k}\right)f_F(\omega _k\mu +Vnz)\right]\frac{d^3\text{k}}{2\mathrm{\Omega }_{BZ}}}`$ (8) $`n_2=n(1\tau )={\displaystyle \left[\left(1+\frac{\mathrm{\Delta }}{\omega _k}\right)f_F(\omega _k\mu +Vnz)+\left(1\frac{\mathrm{\Delta }}{\omega _k}\right)f_F(\omega _k\mu +Vnz)\right]\frac{d^3\text{k}}{2\mathrm{\Omega }_{BZ}}},`$ (9) where $`f_F(y)=1/\left(e^{y/T}+1\right)`$ is the Fermi distribution function, and $`\mathrm{\Omega }_{BZ}`$ is the volume of the first Brillouin zone. Summing up and subtracting two equations for $`n_1`$ and $`n_2`$, we get the resulting system of equations for $`n`$ and $`\mu `$: $$\{\begin{array}{ccc}1=Vz\frac{1}{\omega _k}\left[f_F(\omega _k\mu +Vnz)f_F(\omega _k\mu +Vnz)\right]\frac{d^3𝐤}{2\mathrm{\Omega }_{BZ}}& & \\ & & \\ n=\left[f_F(\omega _k\mu +Vnz)+f_F(\omega _k\mu +Vnz)\right]\frac{d^3𝐤}{2\mathrm{\Omega }_{BZ}}.& & \end{array}$$ (10) For low temperatures $`T0`$ and $`n\frac{1}{2}`$ it is reasonable to assume that $`\mu Vnz`$ is negative. Hence $`f_F(\omega _k\mu +Vnz)=0`$ and $`f_F(\omega _k\mu +Vnz)=\theta (\omega _k\mu +Vnz)`$ is the step function. It is easy to see that for $`n=\frac{1}{2}`$ the system of equations (9) yields identical results for all $`\mathrm{\Delta }\mu Vnz\mathrm{\Delta }`$. From this point of view, $`n=1/2`$ is a point of indifferent equilibrium. For infinitely small deviations from $`n=1/2`$, that is, for densities $`n=1/20`$, the chemical potential should be defined as $`\mu =\mathrm{\Delta }+Vz/2=Vz/2(1\tau )`$. If we consider a strong coupling case $`V2t`$ and assume a constant density of states inside the band, then for a simple cubic lattice we have $`\tau =1\frac{2W^2}{3V^2z^2}`$, and hence $$\mu =\frac{W^2}{3Vz},$$ (11) where $`W=2zt`$ is the bandwidth. Note that for the density $`n=1/2`$ a charge-ordering gap $`\mathrm{\Delta }`$ appears for an arbitrary interaction strength $`V`$. This is due to the existence of nesting in our simple model. In the weak coupling case $`V2t`$ and with the perfect nesting, we have $`\mathrm{\Delta }W\mathrm{exp}\left\{\frac{W}{Vz}\right\}`$ and $`\tau `$ is exponentially small. For $`zVW`$ or accordingly for $`V2t`$: $`\mathrm{\Delta }Vz/2`$ and $`\tau 1`$. As mentioned above, for a general form of electron dispersion without nesting the charge ordering exists only if the interaction strength $`V`$ exceeds certain critical value of the order of bandwidth $`W`$ . Further on, we restrict ourselves only to the physically more instructive case of strong coupling $`V2t`$. Now let us consider the case $`n=1/2\delta `$, where $`\delta 1`$ is a deviation from density $`n=1/2`$. For this case $`\mu =\mu (\delta ,\tau )`$, and we have two coupled equations for $`\mu `$ and $`\tau `$. As a result, $$\mu (\delta )Vnz(1\tau )\frac{4W^2}{Vz}\delta ^2\frac{W^2}{3Vz}+\frac{4W^2}{3Vz}\delta +O(\delta ^2).$$ (12) Correspondingly, the energy of a charge ordered state is as follows $$E_{CO}(\delta )=E_{CO}(0)\frac{W^2}{3Vz}\delta \frac{2W^2}{3Vz}\delta ^2+O(\delta ^3),$$ (13) where $`E_{CO}(0)=\frac{W^2}{6Vz}`$ is the energy precisely for density $`n=1/2`$ and $`|E_{CO}(0)|W`$. At the same time, the charge-ordering gap $`\mathrm{\Delta }`$ is given by: $$\mathrm{\Delta }\frac{Vz}{2}\left[12\delta \frac{2W^2}{3V^2z^2}(1+4\delta )\right].$$ (14) Actually, the dependence of the chemical potential $`\mu `$ and the total energy $`E`$ on $`\delta `$, Eqs. (11), (12), stems from this linear decrease of energy gap $`\mathrm{\Delta }`$ with deviation from half-filling. For $`n>1/2`$ the energy of charge ordered state starts to increase rapidly due to the large contribution from Coulomb repulsion (the upper Verwey band is partially filled for $`n>1/2`$). Contrary to the case $`n=1/2`$, for $`n>1/2`$ each extra electron put into the checkerboard CO state necessarily has occupied nearest neighbor sites, increasing the total energy on $`Vz|\delta |`$. As a result, we have for $`|\delta |=n1/2>0`$ $$E_{CO}(\delta )=E_{CO}(0)+\left(Vz\frac{W^2}{3Vz}\right)|\delta |\frac{2W^2}{3Vz}\delta ^2+O(\delta ^3).$$ (15) Accordingly, the chemical potential has the form $$\mu (\delta )=Vz\frac{W^2}{3Vz}\frac{4W^2}{3Vz}|\delta |+o(\delta ^2).$$ (16) It undergoes a jump equal to $`Vz`$ for $`\tau 1`$. Note that the gap $`\mathrm{\Delta }`$ is symmetric for $`n>1/2`$ and is given by $$\mathrm{\Delta }\frac{Vz}{2}\left[12|\delta |\frac{2W^2}{3V^2z^2}(1+4|\delta |)\right].$$ We could make the whole picture symmetric with respect to $`n=1/2`$ by shifting all the one-electron energy levels and the chemical potential by $`Vz/2`$, i.e., defining $`\mu ^{}=\mu Vz/2`$. This change of variables, of course, would not modify our conclusions. ## III Phase separation The most remarkable implication of (11)-(15) is that the compressibility $`\kappa `$ of the homogeneous charge ordered system is negative for densities different from $`1/2`$, $$\frac{1}{\kappa }\frac{d\mu }{dn}=\frac{d\mu }{d\delta }=\frac{d^2E}{d\delta ^2}=\frac{4W^2}{3Vz}<0,$$ (17) where $`\delta =1/2n`$. This is the manifestation of the tendency toward phase separation characteristic of the charge ordered system with $`\delta 0`$. The presence of a kink in the $`E_{CO}(\delta )`$ (cf. Eqs. (12), (14)) implies that one of the states, into which the system might separate, would correspond to the checkerboard CO state with $`n=1/2`$, whereas the other would have a certain density $`n^{}`$ smaller or larger than 1/2. This conclusion resembles that of (see also ), although the detailed physical mechanism is different. The possibility of phase separation in the model (1) away from half-filling was also reported earlier in for the infinite-dimensional case. Below we concentrate our attention on the situation with $`n<1/2`$ (underdoped manganites); the case $`n>1/2`$ apparently has certain special properties — the existence of stripe phases etc. , the detailed origin of which is not yet clear. It easy to understand the physics of phase separation in our case. As follows from (13), the CO gap decreases linearly with the deviation from the half-filling. Correspondingly, the energy of the homogeneous CO state rapidly increases, and it is more favorable to “extract” extra holes from the CO state, putting them into one part of the sample, while creating the “pure” checkerboard CO state in the other part of it. The energy loss due to such redistribution of holes is overcompensated by the gain provided by the better charge ordering. However, the hole-rich regions would not be completely “empty,” like pores (clusters of vacancies) in crystals: we can gain an extra energy by “dissolving” in them a certain amount of electrons. By doing this we decrease the band energy of the electrons due to their delocalization. Thus, this second phase would be a metallic one. The simplest state of this kind is a homogeneous metal with the electron concentration $`n_m`$. This concentration, as well as the relative volume of the metallic and CO phases, can be easily calculated by minimizing the total energy of the system. The energy of the metallic part of the sample $`E_m`$ is given by $$E_m=tzn_m+ct(n_m)^{5/3}+V(n_m)^2$$ (18) where $`c`$ is some constant. Minimizing (17) with respect to $`n_m`$, we find the equilibrium electron density in the metallic phase. For the strong coupling $`V>zt`$, we get $$n_{m0}=tz/2V$$ (19) Thus, in this simple treatment, the system with $`n_{m0}<n<1/2`$ would undergo phase separation into the CO phase with $`n=1/2`$ and the metallic phase with $`n=n_{m0}`$. Relative volumes $`v_m`$ and $`v_{CO}`$ of these phases for arbitrary $`n`$ can be found from the Maxwell construction: $$v_m/v_{CO}=(1/2n)/(nn_{m0}),$$ (20) from which we find that the metallic phase occupies the part $`v_m`$ of the total volume $`v`$ given by the relationship $$v_m/v=(1/2n)/(1/2n_{m0}),$$ (21) The metallic phase would occupy the whole sample when the total electron density $`n`$ is less than $`n_{m0}`$. ## IV An example: the phase separated state with metallic droplets As we argued above, the system with the short-range repulsion (1) is unstable toward phase separation for $`n`$ close but different from 1/2. The long-range Coulomb forces would however prevent the full phase separation into large regions containing all extra holes and the pure $`n=1/2`$ charge ordered region. We can avoid this energy loss by forming, instead of one big metallic phase with many electrons, the finite metallic clusters with smaller number of them. The limiting case would be a set of spherical droplets, each containing one electron. This state is similar to magnetic polarons (“ferrons”) considered in the problem of phase separation in doped magnetic insulators . We present below the estimate for the characteristic parameters of these droplets. The main aim of this treatment is to demonstrate that the state constructed in such a way will have the energy lower than the energy of the homogeneous state, even if we treat these droplets rather crudely and do not optimize all their properties. In particular, we will make the simplest assumption that the droplets have sharp boundaries and that the charge ordered state outside these droplets is not modified in their vicinity. This state can be treated as a variational one: if we optimize the structure of the droplet boundary, its energy would only decrease. The energy (per unit volume) of the droplet state with the concentration $`n_d`$ of droplets can be written in total analogy with the ferron energy in the double exchange model (see ). This yields $$E_{droplet}=tn_d\left(z\frac{\pi ^2a^2}{R^2}\right)\frac{W^2}{6Vz}\left[1n_d\frac{4}{3}\pi \left(\frac{R}{a}\right)^3\right].$$ (22) Here, $`a`$ is the lattice constant and $`R`$ is the droplet radius. The first term in (21) corresponds to the gain in kinetic energy of electron delocalization inside the metallic droplets, and the second term describes the charge ordering energy in the remaining insulating part of the sample. Actually, one should include the surface energy contribution to the total energy of the droplet. The surface energy should be of the order of $`W^2R^2/V`$. For large droplets, this contribution is small compared to the term $`R^3`$ in (21); it would also be reduced for a “soft” droplet boundary. One can show that even in the worst case of a small droplet (of the order of a few lattice constants) with the sharp boundary, this contribution would not exceed about 20 percent of the bulk contribution. That is why we will ignore this term below. Minimization of the energy in (21) with respect to $`R`$ gives $$\frac{R}{a}\left(\frac{V}{t}\right)^{1/5}.$$ (23) The critical concentration $`n_{dc}`$ corresponds to the configuration where metallic droplets start to overlap, i.e. where the volume of the CO phase ( the second term in (21)) tends to zero. Hence, $$n_{dc}=\frac{3}{4\pi }\left(\frac{a}{R}\right)^3\left(\frac{t}{V}\right)^{3/5}.$$ (24) By comparing (12) with (21), (22), we see that for the deviations from the half-filling $`0<\delta \delta _c=1/2n_{dc}`$ the energy of the phase separated state is always lower than the energy of the homogeneous charge ordered state. The energy of the droplet state (21) with the radius given by (22) is also lower than the energy of the fully phase separated state, obtained by the Maxwell construction from the homogeneous metallic state (17). Correspondingly, the critical concentration $`n_{dc}`$ (23) is larger than $`n_{m0}`$ (18). There is no contradiction here: the droplet state, which we constructed has electrons confined in spheres of radius $`R`$, and even when these droplets start to overlap at $`n=n_{dc}`$, occupying the whole sample, the electrons in this state, by construction, are still confined within their own spheres and avoid each other. In other words, in our droplet state certain degree of charge-ordering correlations is still present, decreasing the repulsion and hence the total energy. Thus, the energy of a phase separated state with the droplets corresponds to the global minima of the energy for all $`0<\delta \delta _c`$. This justifies our conclusion about phase separation into charge ordered state with $`n=1/2`$ and a metallic state with small spherical droplets. The situation here resembles that of partially filled strongly interacting Hubbard model, with the CO state corresponding to an antiferromagnetic state of the latter and with the nearest-neighbor interaction $`V`$ playing the role of the Hubbard’s $`U`$. In both cases, the kinetic energy of doped carriers tends to destroy this “antiferro” or charge ordering, first “spoiling” it in their vicinity and finally leading to the formation of the metallic state (Nagaoka ferromagnetism). In the Hubbard model, we also face the situation with phase separation at a small enough doping . Note also that for $`n>1/2`$ the compressibility of the charge- ordered state is again negative $`\frac{1}{\kappa }=\frac{d^2E}{d\delta ^2}=\frac{4W^2}{3Vz}<0`$ and has the same value as for the case $`n<1/2`$. As a result, it is more favorable again to create a phase-separated state for these densities. However, as it was already mentioned, the nature of the second phase with $`n>1/2`$ is not quite clear at present, and therefore we do not consider this case here. ## V An extended model Now we can extend the model discussed in the previous sections by taking into account the essential magnetic interactions. In manganites, besides the conduction electrons in $`e_g`$ bands, there exist also practically localized $`t_{2g}`$ electrons, which we now include to our consideration. The corresponding Hamiltonian has the form $$\widehat{H}=t\underset{<i,j>,\sigma }{}c_{i\sigma }^+c_{j\sigma }+V\underset{<i,j>}{}n_in_jJ_H\underset{i}{}𝐒_i\sigma _i+J\underset{<i,j>}{}𝐒_i𝐒_j\mu \underset{i}{}n_i,$$ (25) In comparison to (1), the additional terms here correspond to the strong Hund-rule onsite coupling $`J_H`$ between localized spins $`𝐒`$ and the spins of conduction electrons $`\sigma `$, and a relatively weak Heisenberg antiferromagnetic (AFM) exchange $`J`$ between neighboring local spins. In real manganites, the AFM ordering of the CE type in the CO phase is determined not only by the exchange of localized $`t_{2g}`$ electrons but to a large extent by the charge- and orbitally-ordered $`e_g`$ electrons themselves. For simplicity, we ignore this factor here and assume the superexchange interaction to be the same both in the CO and in the metallic phases. It is physically reasonable to consider this model in the limit $`J_HS>V>W>JS^2.`$ In the absence of the Coulomb term, this is exactly the conventional double exchange model (see e.g. ). Note that the absence of doubly occupied sites in (20) is guaranteed by the large Hund’s term. It also favors the metallicity in the system, since the effective bandwidth in our problem depends upon the magnetic order. Therefore, the estimate for the critical concentration changes here in comparison to (23). Similar to the metallic droplets will be ferromagnetic (FM) due to the double exchange. The energy of one such droplet has the form $`E`$ $`=`$ $`t\left(z{\displaystyle \frac{\pi ^2a^2}{R^2}}\right){\displaystyle \frac{W^2}{6Vz}}\left[1{\displaystyle \frac{4}{3}}\pi \left({\displaystyle \frac{R}{a}}\right)^3\right]+`$ (26) $`+`$ $`zJS^2{\displaystyle \frac{4}{3}}\pi \left({\displaystyle \frac{R}{a}}\right)^3zJS^2\left[1{\displaystyle \frac{4}{3}}\pi \left({\displaystyle \frac{R}{a}}\right)^3\right].`$ (27) The last two terms in (25) describe respectively the loss in the energy of the Heisenberg AFM exchange inside the FM metallic droplets and the gain of this energy in the AFM insulating part of the sample. The minimization with respect to the droplet radius (as in (21)) yields $$\frac{R}{a}\left(\frac{t}{V}+\frac{JS^2}{t}\right)^{1/5}.$$ (28) Note that at $`t/VJS^2/t`$, formula (26) gives just the same estimate for the radius of FM metallic droplet $`(R/a)(t/JS^2)^{1/5}`$ as in . In the opposite limit when $`(t/V)JS^2/t`$, we reproduce the same result $`(R/a)(V/t)^{1/5}`$ as in (22). Finally, critical concentration $`n_c`$ is estimated as follows $$n_c\left(\frac{t}{V}+\frac{JS^2}{t}\right)^{3/5}.$$ (29) As a result, taking into account also the tendency to the phase separation at very small values of $`n`$ we come to the following phase diagram for the extended model (cf. : 1. At $`0<n<\left(\frac{JS^2}{t}\right)^{3/5}`$, it corresponds to the phase separation into a FM metal with $`n=n^{}>0`$ embedded in the AFM insulating matrix ($`n=0`$). To minimize the Coulomb energy, it may be again favorable to split this metallic region into droplets with the concentration $`n^{}`$ and an average radius given by Eq. (26) with $`t/V=0`$, each containing one electron and kept apart from one another. 2. At $`\left(\frac{JS^2}{t}\right)^{3/5}<n<\left(\frac{t}{V}+\frac{JS^2}{t}\right)^{3/5}<1/2`$, the system is a FM metal. Of course, we need a window of parameters to satisfy the inequality in the right-hand side. In actual manganites where $`t/V1/2÷1/3`$ and $`J/t1/3`$, these conditions upon $`n`$ are not necessarily satisfied. Experiments suggest that this window is present for La<sub>1-x</sub>Ca<sub>x</sub>MnO<sub>3</sub>, but it is definitely absent for Pr<sub>1-x</sub>Ca<sub>x</sub>MnO<sub>3</sub> ; 3. Finally, at $`\left(\frac{t}{V}+\frac{JS^2}{t}\right)^{3/5}<n<\frac{1}{2}`$, we have the phase separation in the form of FM metallic droplets inside the AFM charge ordered matrix. This phase diagram is in a good qualitative agreement with many available experimental results for real manganites , in particular with the observation of the small-scale phase separation close to 0.5 doping . Note also that our phase diagram has certain similarities with the phase diagram obtained in for the problem of spontaneous ferromagnetism in doped excitonic insulators. ## VI Conclusion Summarizing, we have shown that the narrow-band system, which has the checkerboard charge ordering at $`n=1/2`$ (corresponding to the doping $`x=0.5`$) is unstable toward phase separation away from half-filling ($`n1/2`$). It separates into the regions with the ideal CO ($`n=1/2`$) and the other regions, in which extra electrons or holes are trapped. The simplest form of these metallic regions could be spherical metallic droplets embedded into the CO insulating matrix. Simple considerations allow us to estimate the size of these droplets and the critical concentration, or doping $`x_c=1/2\delta _c`$, at which the metallic phase would occupy the whole sample and the CO phase would disappear. The account of the magnetic interactions does not change these conclusions but somewhat modifies the characteristic parameters of the metallic droplets. The long-range Coulomb interaction may also modify the results, but we do not expect any qualitative changes. For the realistic values of parameters, the size of metallic droplets is still microscopic (about 10–30 $`\AA `$), and the excess charge in them will be rather small. The obtained picture corresponds rather well to the known properties of 3D and layered manganites close to (less than) half doping, $`x1/2`$. Percolation picture of transport properties emerging from this treatment is confirmed by the results reported in ; moreover the coexistence of ferromagnetic reflections and those of the CE type magnetic structure typical of the CO state at $`x=0.5`$ were observed by the neutron scattering . Thus, the general behavior of underdoped manganites ($`x0.5`$) is in a good qualitative agreement with our results. Our treatment leads to the same tendency to phase separation (instability of the homogeneous CO phase) also for overdoped regime, $`x>0.5`$. What would be the second phase in this case, is not yet clear. Therefore we did not concentrate our attention on such a situation. Our treatment is applicable also to other systems with the charge ordering, such as cobaltites and nickelates . It would be interesting to study them for charge carrier concentrations different from the commensurate “checkerboard” one A number of important problems still remain unsolved (the origin of the “in-phase” ordering in perovskite manganites in $`c`$-direction, the detailed description of inhomogeneous states in overdoped regime $`n>1/2`$, the behavior at finite temperatures). Nevertheless, in spite of the introduced simplifications, our model seems to capture the essential physics underlying the interplay between phase separation and charge ordering in transition metal oxides. ## Acknowledgments We are grateful to N.M. Plakida and M.S. Mar’enko for stimulating discussions. D.Kh. expresses gratitude to S.-W. Cheong and Y. Moritomo for the discussions of the experimental aspects of the problem. The work was supported by INTAS (grants 97–0963 and 97–11954) and by the Russian-Dutch Program for Scientific Cooperation funded by the Netherlands Organization for Scientific Research (NWO). M.Yu.K. acknowledges the support of the Russian President Program (grant 96–15–9694). The work of D.Kh. was also supported by the Netherlands Foundation for the Fundamental Research of Matter (FOM) and by the European network OXSEN.
warning/0001/nucl-th0001064.html
ar5iv
text
# 𝑇-dependence of pseudoscalar and scalar correlations ## I Introduction Numerical simulations of $`2`$-light-flavour lattice-QCD, and the study of models that accurately describe dynamical chiral symmetry breaking and the $`\pi `$-$`\rho `$ mass-difference at $`T=0`$ both indicate that a quark-gluon plasma (QGP) is reached via a second order transition at a critical temperature $`T_c0.15`$GeV. This QGP existed as a stage in the early evolution of the universe and its terrestrial recreation is a primary goal of current-generation ultrarelativistic heavy-ion experiments. The plasma is characterised by the free propagation of quarks and gluons over distances $`10`$-times larger than the proton. However, its formation can only be observed indirectly by searching for a modification of particle yields and/or hadron properties in the debris of the collisions. The plasma phase is also characterised by chiral symmetry restoration. Hence, since the properties of the pion (mass, decay constant, other vertex residues, etc.) are tied to the dynamical breaking of chiral symmetry, an elucidation of the $`T`$-dependence of these properties is important; particularly since a prodigious number of pions is produced in heavy ion collisions. Also important is understanding the $`T`$-dependence of the properties of the scalar analogues and chiral partners of the pion in the strong interaction spectrum. For example, should the mass of a putative light isoscalar-scalar meson fall below $`2m_\pi `$, the strong decay into a two pion final state can no longer provide its dominant decay mode. In this case electroweak processes will be the only open decay channels below $`T_c`$ and the state will appear as a narrow resonance. Analogous statements are true of isovector-scalar mesons. The Dyson-Schwinger equations (DSEs) provide a nonperturbative, continuum framework for analysing quantum field theories and the class of rainbow-ladder truncation DSE models yields a qualitative understanding of the thermodynamic properties of the QGP phase transition at $`T0`$ (and also at $`\mu 0`$). For example, using a simple element of this class the pressure’s slow approach to its ultrarelativistic limit, which is observed in numerical simulations of lattice-QCD, could be attributed to a persistence into the QGP of nonperturbative effects in the quark’s vector self energy.<sup>*</sup><sup>*</sup>*$`T=0`$ DSE studies characteristically yield momentum-dependent vector and scalar quark self energies that remain large until $`k^2=1`$GeV<sup>2</sup>; e.g., Fig. 8 of Ref. and Ref. . These features are also observed in lattice simulations. Also, using a more sophisticated model, the present incompatibility between lattice estimates of the critical exponents was identified as likely an artefact of working too far from the chiral limit. The calculation of the $`T`$-dependence of hadron properties has hitherto used only simple members of this class. Herein we employ a version that provides for renormalisability and the correct one-loop renormalisation group evolution of scale-dependent matrix elements, and focus on scalar and pseudoscalar properties. Mesons appear as simple poles in $`3`$-point vertices. Importantly, however, these vertices also provide information about the persistence of correlations away from the bound state pole; e.g., Refs. , which can be useful in studying the $`T`$-evolution of a system with deconfinement. Consequently our starting point is the inhomogeneous Bethe-Salpeter equation (BSE). We then employ the homogeneous equation when appropriate and useful. In Sec. II we describe the $`T0`$ quark DSE and inhomogeneous BSEs, and introduce our model. This also serves to make clear our notation. Section III reports our results for the $`T`$-dependence of scalar and pseudoscalar correlations while Sec. IV is a brief recapitulation and epilogue. An appendix contains selected formulae. ## II Dyson-Schwinger and Bethe-Salpeter equations The renormalised quark DSE is $`S^1(p_{\omega _k})`$ $`:=`$ $`i\stackrel{}{\gamma }\stackrel{}{p}A(p_{\omega _k})+i\gamma _4\omega _kC(p_{\omega _k})+B(p_{\omega _k})`$ (1) $`=`$ $`Z_2^Ai\stackrel{}{\gamma }\stackrel{}{p}+Z_2(i\gamma _4\omega _k+m_{\mathrm{bm}})+\mathrm{\Sigma }^{}(p_{\omega _k}),`$ (2) where $`p_{\omega _k}:=(\stackrel{}{p},\omega _k)`$ with $`\omega _k=(2k+1)\pi T`$ the fermion Matsubara frequency, and $`m_{\mathrm{bm}}`$ is the Lagrangian current-quark bare mass. The regularised self energy is $$\mathrm{\Sigma }^{}(p_{\omega _k})=i\stackrel{}{\gamma }\stackrel{}{p}\mathrm{\Sigma }_A^{}(p_{\omega _k})+i\gamma _4\omega _k\mathrm{\Sigma }_C^{}(p_{\omega _k})+\mathrm{\Sigma }_B^{}(p_{\omega _k}),$$ (3) $`\mathrm{\Sigma }_{}^{}(p_{\omega _k})`$ $`=`$ $`{\displaystyle _{l,q}^{\overline{\mathrm{\Lambda }}}}{\displaystyle \frac{4}{3}}g^2D_{\mu \nu }(\stackrel{}{p}\stackrel{}{q},\omega _k\omega _l)`$ (5) $`\times {\displaystyle \frac{1}{4}}\mathrm{tr}\left[𝒫_{}\gamma _\mu S(q_{\omega _l})\mathrm{\Gamma }_\nu (q_{\omega _l};p_{\omega _k})\right],`$ where: $`=A,B,C`$; $`A,B,C`$ are functions of $`(|\stackrel{}{p}|^2,\omega _k^2)`$; $$𝒫_A:=(Z_1^A/|\stackrel{}{p}|^2)i\stackrel{}{\gamma }\stackrel{}{p},𝒫_B:=Z_1,𝒫_C:=(Z_1/\omega _k)i\gamma _4;$$ (6) and $`_{l,q}^{\overline{\mathrm{\Lambda }}}:=T_{l=\mathrm{}}^{\mathrm{}}^{\overline{\mathrm{\Lambda }}}d^3q/(2\pi )^3`$, with $`^{\overline{\mathrm{\Lambda }}}`$ representing a translationally invariant regularisation of the integral and $`\overline{\mathrm{\Lambda }}`$ the regularisation mass-scale. The renormalised self energies are $$\begin{array}{ccc}\hfill (p_{\omega _k};\zeta )& =& \xi _{}+\mathrm{\Sigma }_{}^{}(p_{\omega _k};\overline{\mathrm{\Lambda }})\mathrm{\Sigma }_{}^{}(\zeta _{\omega _0}^{};\overline{\mathrm{\Lambda }}),\hfill \end{array}$$ (7) $`\zeta `$ is the renormalisation point, $`(\zeta _{\omega _0}^{})^2:=\zeta ^2\omega _0^2`$, $`\xi _A=1=\xi _C`$, and $`\xi _B=m_R(\zeta )`$. ### A The Model $`\mathrm{\Gamma }_\nu (q_{\omega _l};p_{\omega _k})`$ in Eq. (5) is the renormalised dressed-quark-gluon vertex. It is a connected, irreducible $`3`$-point function that should not exhibit light-cone singularities in covariant gauges . A number of Ansätze with this property have been proposed and it has become clear that the judicious use of the rainbow truncation $$\mathrm{\Gamma }_\nu (q_{\omega _l};p_{\omega _k})=\gamma _\nu $$ (8) in Landau gauge provides phenomenologically reliable results so we employ it herein. A mutually consistent constraint is $$Z_1=Z_2\mathrm{and}Z_1^A=Z_2^A.$$ (9) The rainbow truncation is the leading term in a $`1/N_c`$ expansion of $`\mathrm{\Gamma }_\nu (q_{\omega _l};p_{\omega _k})`$. $`D_{\mu \nu }(p_{\mathrm{\Omega }_k})`$ is the renormalised dressed-gluon propagator $$g^2D_{\mu \nu }(p_{\mathrm{\Omega }_k})=P_{\mu \nu }^L(p_{\mathrm{\Omega }_k})\mathrm{\Delta }_F(p_{\mathrm{\Omega }_k})+P_{\mu \nu }^T(p_{\mathrm{\Omega }_k})\mathrm{\Delta }_G(p_{\mathrm{\Omega }_k}),$$ (10) $`P_{\mu \nu }^T(p_{\mathrm{\Omega }_k})`$ $`:=`$ $`\{\begin{array}{cc}0,\hfill & \mu \mathrm{and}/\mathrm{or}\nu =4,\hfill \\ \delta _{ij}{\displaystyle \frac{p_ip_j}{p^2}},\hfill & \mu ,\nu =i,j=1,2,3,\hfill \end{array}`$ (13) $`P_{\mu \nu }^L(p_{\mathrm{\Omega }_k})`$ $`+`$ $`P_{\mu \nu }^T(p_{\mathrm{\Omega }_k})=\delta _{\mu \nu }{\displaystyle \frac{p_\mu p_\nu }{p^2}},`$ (14) $`\mu ,\nu =1,\mathrm{},4`$, and $`\mathrm{\Omega }_k:=2\pi kT`$. A Debye mass for the gluon appears as a $`T`$-dependent contribution to $`\mathrm{\Delta }_F`$. The ultraviolet behaviour of the kernel in Eq. (5) is fixed by perturbative QCD because the DSEs yield perturbation theory in the weak coupling limit. Our model is defined by specifying a form for the kernel’s infrared behaviour: $`\mathrm{\Delta }_F(p_{\mathrm{\Omega }_k})`$ $`=`$ $`𝒟(p_{\mathrm{\Omega }_k};m_g),\mathrm{\Delta }_G(p_{\mathrm{\Omega }_k})=𝒟(p_{\mathrm{\Omega }_k};0),`$ (15) $`𝒟(p_{\mathrm{\Omega }_k};m_g)`$ $`:=`$ $`2\pi ^2D{\displaystyle \frac{2\pi }{T}}\delta _{0k}\delta ^3(\stackrel{}{p})+𝒟_\mathrm{M}(p_{\mathrm{\Omega }_k};m_g),`$ (16) where $`D=(0.881\mathrm{GeV})^2`$ is a mass-scale parameter and $`𝒟_\mathrm{M}(p_{\mathrm{\Omega }_k};m_g)={\displaystyle \frac{4\pi ^2}{\omega ^6}}Ds_{\mathrm{\Omega }_k}\mathrm{e}^{s_{\mathrm{\Omega }_k}/\omega ^2}`$ (18) $`+{\displaystyle \frac{8\pi ^2\gamma _m}{\mathrm{ln}\left[\tau +\left(1+s_{\mathrm{\Omega }_k}/\mathrm{\Lambda }_{\mathrm{QCD}}^2\right)^2\right]}}{\displaystyle \frac{1\mathrm{e}^{s_{\mathrm{\Omega }_k}/(4m_t^2)}}{s_{\mathrm{\Omega }_k}}},`$ with $`s_{\mathrm{\Omega }_k}=p_{\mathrm{\Omega }_k}^2+m_g^2`$, $`\tau =e^21`$, $`\omega =1.2m_t`$, $`m_t=0.5`$GeV, $`\gamma _m=12/25`$, $`m_g^2=(16/5)\pi ^2T^2`$, and $`\mathrm{\Lambda }_{\mathrm{QCD}}^{N_f=4}=0.234`$GeV. This model, which is motivated by Refs. , incorporates the one-loop logarithmic suppression identified in perturbative calculations. Its parameters were fixed at $`T=0`$ by fitting a range of $`\pi `$ and $`K`$ meson properties: a renormalisation point invariant light current-quark mass $`\widehat{m}_{u,d}=5.7`$MeV, corresponding to $`m_R(1\mathrm{GeV})=4.8`$MeV, gives $`m_\pi =0.14`$GeV, $`f_\pi =0.092`$MeV. The model exhibits a second order chiral symmetry restoring transition at $$T_c=0.15\mathrm{GeV}.$$ (19) ### B Pseudoscalar Channel Ward-Takahashi identities relate the $`3`$-point vector and axial-vector vertices to the dressed-quark propagator. Hence, once a truncation of the kernel in the quark DSE has been selected, requiring the preservation of these identities constrains the kernel in the BSE. This is explored in Ref. , where a systematic procedure for constructing the kernels is introduced that ensures the order-by-order preservation of these identities. Using this procedure the inhomogeneous BSE for the zeroth Matsubara mode of the isovector $`0^+`$ vertex consistent with the rainbow truncation of Eq. (2) is $`\mathrm{\Gamma }_{\mathrm{ps}}^i(p_{\omega _k};P_0;\zeta )=Z_4{\displaystyle \frac{1}{2}}\tau ^i\gamma _5`$ (20) $`{\displaystyle _{l,q}^{\overline{\mathrm{\Lambda }}}}{\displaystyle \frac{4}{3}}g^2D_{\mu \nu }(p_{\omega _k}q_{\omega _l})`$ (21) $`\times `$ $`\gamma _\mu S(q_{\omega _l}^+)\mathrm{\Gamma }_{\mathrm{ps}}^i(q_{\omega _l};P_0;\zeta )S(q_{\omega _l}^{})\gamma _\nu ,`$ (22) where $`\{\tau ^i`$, $`i=1,2,3`$} are the Pauli matrices, $`q_{\omega _l}^\pm =q_{\omega _l}\pm P_0/2`$, $`P_0=(\stackrel{}{P},0)`$, and $`Z_4=Z_4(\zeta ,\overline{\mathrm{\Lambda }})`$ is the mass renormalisation constant: $$m_R(\zeta )\mathrm{\Gamma }_{\mathrm{ps}}^i(p_{\omega _k};P_{\mathrm{\Omega }_n};\zeta )$$ (23) is renormalisation point independent. Equation (22) is a dressed-ladder BSE. At $`T=0`$ its solution exhibits poles, and their positions and residues provide a good description of light vector and flavour-nonsinglet pseudoscalar mesons when $`S`$ is obtained from the rainbow quark DSE. This truncation is reliable because of cancellations between vertex corrections and crossed-box contributions at each higher order in the quark-antiquark scattering kernel. The solution of Eq. (22) has the form (hereafter the $`\zeta `$-dependence is often implicit) $`\mathrm{\Gamma }_{\mathrm{ps}}^i(p_{\omega _k};\stackrel{}{P})=`$ (26) $`{\displaystyle \frac{1}{2}}\tau ^i\gamma _5[iE_{\mathrm{ps}}(p_{\omega _k};\stackrel{}{P})+\stackrel{}{\gamma }\stackrel{}{P}F_{\mathrm{ps}}(p_{\omega _k};\stackrel{}{P})`$ $`+\stackrel{}{\gamma }\stackrel{}{p}\stackrel{}{p}\stackrel{}{P}G_{\mathrm{ps}}^{}(p_{\omega _k};\stackrel{}{P})+\gamma _4\omega _k\stackrel{}{p}\stackrel{}{P}G_{\mathrm{ps}}^{}(p_{\omega _k};\stackrel{}{P})],`$ where we have neglected terms involving $`\sigma _{\mu \nu }`$-like contributions, which play a negligible role at $`T=0`$. The scalar functions in Eq. (26) exhibit a simple pole at $`\stackrel{}{P}^2+m_\pi ^2=0`$ so that $$\mathrm{\Gamma }_{\mathrm{ps}}^i(p_{\omega _k};\stackrel{}{P})=\frac{r_\pi (\zeta )}{\stackrel{}{P}^2+m_\pi ^2}\mathrm{\Gamma }_\pi ^i(p_{\omega _k};\stackrel{}{P})+\mathrm{regular},$$ (27) where “regular” means terms regular at this pole and $`\mathrm{\Gamma }_\pi ^i(p_{\omega _k};\stackrel{}{P})`$ is the canonically normalised, bound state pion Bethe-Salpeter amplitude: $`2\delta ^{ij}\stackrel{}{P}=\mathrm{tr}{\displaystyle _{l,q}^{\overline{\mathrm{\Lambda }}}}\{\mathrm{\Gamma }_\pi ^i(q_{\omega _l};\stackrel{}{P}){\displaystyle \frac{S(q_{\omega _l}^+)}{\stackrel{}{P}}}\mathrm{\Gamma }_\pi ^j(q_{\omega _l};\stackrel{}{P})S(q_{\omega _l}^{})`$ (29) $`+\mathrm{\Gamma }_\pi ^i(q_{\omega _l};\stackrel{}{P})S(q_{\omega _l}^+)\mathrm{\Gamma }_\pi ^j(q_{\omega _l};\stackrel{}{P}){\displaystyle \frac{S(q_{\omega _l}^{})}{\stackrel{}{P}}}\left\}\right|_{\stackrel{}{P}^2=m_\pi ^2},`$ with the trace over colour, Dirac and isospin indices, and the residue is $$\delta ^{ij}ir_\pi =Z_4\mathrm{tr}_{l,q}^{\overline{\mathrm{\Lambda }}}\frac{1}{2}\tau ^i\gamma _5\chi _\pi ^j(q_{\omega _l};\stackrel{}{P}),$$ (30) where $`\chi _\pi (q_{\omega _l};\stackrel{}{P}):=S(q_{\omega _l}^+)\mathrm{\Gamma }_\pi (q_{\omega _l};\stackrel{}{P})S(q_{\omega _l}^{})`$ is the unamputated Bethe-Salpeter wave function. Substituting Eq. (27) into Eq. (22) and equating pole residues yields the homogeneous pion BSE, which provides the simplest way to obtain the bound state amplitude. At $`T=0`$, $`r_\pi (\zeta )`$ is the gauge-invariant pseudoscalar projection of the pion Bethe-Salpeter wave function at the origin in configuration space; i.e, it is a field theoretical analogue of the “wave function at the origin,” which describes the decay of bound states in quantum mechanics. The residue of the pion pole in the axial-vector vertex is the pion decay constant: $$\delta ^{ij}\stackrel{}{P}f_\pi =Z_2^A\mathrm{tr}_{l,q}^{\overline{\mathrm{\Lambda }}}\frac{1}{2}\tau ^i\gamma _5\stackrel{}{\gamma }\chi _\pi ^j(q_{\omega _l};\stackrel{}{P}).$$ (31) It is the gauge-invariant pseudovector projection of Bethe-Salpeter wave function at the origin, which completely determines the strong interaction contribution to the leptonic decay of the pion: $$\mathrm{\Gamma }_{\pi \mathrm{}\nu _{\mathrm{}}}=\frac{1}{4\pi }f_\pi ^2G_F^2|V_{ud}|^2m_\pi m_{\mathrm{}}^2\left(1m_{\mathrm{}}^2/m_\pi ^2\right)^2,$$ (32) $`G_F=1.166\times 10^5`$GeV<sup>-2</sup>, $`|V_{ud}|=0.975`$, and $`m_e=0.511`$MeV, $`m_\mu =0.106`$GeV. It is a model independent consequence of the axial-vector Ward-Takahashi identity that $$f_\pi m_\pi ^2=2m_R(\zeta )r_\pi (\zeta ).$$ (33) In the chiral limit $$\underset{\widehat{m}0}{lim}r_\pi (\zeta )=\frac{1}{f_\pi ^0}\overline{q}q_\zeta ^0,$$ (34) where $`f_\pi ^0`$ is the chiral limit decay constant, and in this model $`f_\pi ^0=0.088\mathrm{GeV},\overline{q}q_{1\mathrm{GeV}^2}^0=(0.235\mathrm{GeV})^3`$ (35) $``$ $`r_\pi ^0(1\mathrm{GeV}^2)=(0.384\mathrm{GeV})^2.`$ (36) ### C Scalar Channel The analogue of Eq. (22) for the $`0^{++}`$ vertex is presented in Eq. (51). However, the combination of rainbow and ladder truncations is not certain to provide a reliable approximation in the scalar sector because here the cancellations described above do not occur. This is entangled with the phenomenological difficulties encountered in understanding the composition of scalar resonances below $`1.4`$GeV. For the isoscalar-scalar vertex the problem is exacerbated by the presence of timelike gluon exchange contributions to the kernel, which are the analogue of those diagrams expected to generate the $`\eta `$-$`\eta ^{}`$ mass splitting in BSE studies. Nevertheless, in the absence of an improved, phenomenologically efficacious kernel we employ Eq. (51) in the expectation that it will provide some qualitatively reliable insight. (This is justified a posteriori.) The scalar functions in Eq. (51) exhibit a simple pole at $`\stackrel{}{P}^2+m_\sigma ^2=0`$, Eqs. (54,56), with residue $$\delta ^{\alpha \beta }r_\sigma =Z_4\mathrm{tr}_{l,q}^{\overline{\mathrm{\Lambda }}}\frac{1}{2}\tau ^\alpha \chi _\sigma ^\beta (q_{\omega _l};\stackrel{}{P}),$$ (37) where $`\chi _\sigma ^\alpha `$ is an obvious analogue of $`\chi _\pi ^i`$. Since a $`VA`$ current cannot connect a $`0^{++}`$ state to the vacuum the scalar meson does not appear as a pole in the vector vertex; i.e, $$\delta ^{\alpha \beta }\stackrel{}{P}f_\sigma =Z_2^A\mathrm{tr}_{l,q}^{\overline{\mathrm{\Lambda }}}\frac{1}{2}\tau ^\alpha \stackrel{}{\gamma }\chi _\sigma ^\beta (q_{\omega _l};\stackrel{}{P})0.$$ (38) The homogeneous equation for the scalar bound state amplitude is obtained from Eqs. (51,54). ### D Two-body Decays The $`\sigma `$ and $`\pi `$ bound state amplitudes along with the dressed-quark propagators are necessary elements in the definition of the impulse approximation to hadronic matrix elements. For example, the isoscalar-scalar-$`\pi \pi `$ coupling is described by the matrix element in Eq. (59). This yields the width $`\mathrm{\Gamma }_{\sigma (\pi \pi )}`$ $`=`$ $`{\displaystyle \frac{3}{2}}g_{\sigma \pi \pi }^2{\displaystyle \frac{\sqrt{14m_\pi ^2/m_\sigma ^2}}{16\pi m_\sigma }},`$ (39) which vanishes if $`m_\sigma <2m_\pi `$. A contemporary analysis of $`\pi \pi `$ data identifies a $`\overline{u}u`$+$`\overline{d}d`$ scalar with $$m_\sigma 0.46\mathrm{GeV},\mathrm{\Gamma }_\sigma 0.220.47\mathrm{GeV},$$ (40) which corresponds to $`g_{\sigma \pi \pi }2.1`$$`3.0`$GeV $`=4.5`$$`6.6m_\sigma `$. We also consider the electromagnetic decay of the neutral pion, for which the dressed-quark-photon vertex is also required in calculating the impulse approximation to the coupling. Quantitatively reliable numerical solutions of the $`T=0`$ vector vertex equation are now available. However, this anomalous coupling is insensitive to details and an accurate result requires only that the dressed vertex satisfy the vector Ward-Takahashi identity. This is similar to the pion form factor: the $`q^2=0`$ value is fixed by current conservation and only the charge radius responds to changes in the vertex, and the $`\gamma ^{}\pi ^0\gamma `$ transition form factor whose value is fixed at the real photon point but whose $`q^2`$-evolution is sensitive to details of the vertex. An efficacious $`T=0`$ vertex Ansatz is given in Eq. (63) and using this in calculating the $`\widehat{m}=0`$ $`\pi ^0\gamma \gamma `$ coupling (Eq. (65) for $`T0`$) one obtains $$g_{\pi ^0\gamma \gamma }^0=\frac{1}{2}$$ (41) independent of the model parameters. Using this chiral limit coupling on-shell $$\mathrm{\Gamma }_{\pi ^0\gamma \gamma }\frac{m_\pi ^3}{16\pi }\frac{\alpha _{\mathrm{em}}^2}{\pi ^2}\left(\frac{g_{\pi ^0\gamma \gamma }^0}{f_\pi }\right)^2=\frac{m_\pi ^3}{64\pi }\left(\frac{\alpha _{\mathrm{em}}}{\pi f_\pi }\right)^2;$$ (42) i.e., $`7.7`$keV cf. the experimental value: $`7.7\pm 0.6`$. ## III Meson Properties Solving the (in)homogeneous BSE at $`T=0`$ is a demanding numerical task because an accurate solution requires a large amount of computer memory and/or time. These problems are exacerbated at $`T0`$ because of the loss of $`O(4)`$ invariance, which is manifest in the separation of the four-momentum into a three-momentum and a Matsubara frequency. Hence in the sum over fermion Matsubara modes we usually limit ourselves to $`l=4,\mathrm{},4`$. At the critical temperature this corresponds to $`\omega _{l=4}>4.0`$GeV, which is greater than the other mass-scales in the problem, and yields results that are numerically accurate to within $`5`$%. Reproducing the $`T=0`$ limit requires many more Matsubara modes, which precludes that as a check of our numerical method. Instead we estimate the error by reducing the number of modes and comparing the results. Our discretisation of the three-momentum grids in the quark DSE and BSE is less seriously limited, and we typically employ $`>1000`$ Gaussian quadrature points. ### A Chiral limit For $`\widehat{m}=0`$ the $`T=0`$ analogues of the inhomogeneous BSEs, Eqs. (22,51), exhibit poles at $$m_\pi =0\mathrm{and}m_\sigma =0.56\mathrm{GeV},$$ (43) with $`m_\sigma =0.59`$GeV at $`\widehat{m}=5.7`$MeV. A low-mass scalar is typical of the rainbow-ladder truncation. However, there is some model sensitivity; e.g., cf. this result with $`m_\sigma =0.59`$GeV in Ref. , $`m_\sigma =0.67`$GeV using the model of Ref. and $`m_\sigma =0.72`$GeV in the separable model of Ref. . These four independent calculations give an average-$`m_\sigma =0.64`$GeV with a standard deviation of $`10`$%. The rainbow-ladder truncation yields degenerate isoscalar and isovector bound states, and ideal flavour mixing in the $`3`$-flavour case. Hence the distribution of mass estimates between that of the isoscalar $`\sigma `$ and isovector $`a_0(980)`$ might be anticipated. In contrast the three comparison studies, Refs. , give $`m_\omega =m_\rho =0.75`$GeV with a standard deviation of $`<2`$%, illustrating the dependability of the truncation in the vector channel. As already remarked, improvements to the kernel are required in the scalar channel. In the isoscalar-scalar channel, because $`\mathrm{\Gamma }_\sigma /m_\sigma `$ is large, it may even be necessary to include couplings to the dominant $`\pi \pi `$ mode, which can be handled perturbatively in the $`\omega `$-$`\rho `$ sector. In the absence of such corrections, the $`\sigma `$ properties elucidated herein are strictly only those of an idealised chiral partner of the $`\pi `$. Hitherto no model bound-state description escapes this caveat. The evolution with $`T`$ of the pole positions in the solution of the inhomogeneous BSEs is illustrated in Fig. 1, from which it is clear that: 1) at the critical temperature, $`T_c`$, we have degenerate, massless pseudoscalar and scalar bound states; and 2) the bound states persist above $`T_c`$, becoming increasingly massive with increasing $`T`$. These features are also observed in numerical simulations of lattice-QCD. We obtain the bound state amplitudes from the homogeneous Bethe-Salpeter equations, which Fig. 1 demonstrates are certain to have a solution. Their $`T`$-evolution is depicted in Fig. 2, which indicates that: 1) in both cases all but the leading Dirac amplitude vanishes above $`T_c`$; and 2) the surviving pseudoscalar amplitude is pointwise identical to the surviving scalar one. These results indicate that the chiral partners are locally identical above $`T_c`$, they do not just have the same mass. It is easy to understand this algebraically. The BSE is a set of coupled homogeneous equations for the Dirac amplitudes. Below $`T_c`$ each of the equations for the subleading Dirac amplitudes has an “inhomogeneity” whose magnitude is determined by $`B_0`$, the scalar piece of the quark self energy which is dynamically generated in the chiral limit. $`B_0`$ vanishes above $`T_c`$ eliminating the inhomogeneity and allowing a trivial, identically zero solution for each of these amplitudes. Additionally, with $`B_00`$ the kernels in the equations for the dominant pseudoscalar and scalar amplitudes are identical, and hence so are the solutions. It follows from these results that the Goldberger-Treiman-like relation $$f_\pi ^0E_\pi (p_{\omega _k};0)=B_0(p_{\omega _k}),$$ (44) is satisfied for all $`T`$ only because both $`f_\pi ^0`$ and $`B_0(p_{\omega _k})`$ are equivalent order parameters for chiral symmetry restoration. This possibility was overlooked in Ref. . Further, above $`T_c`$, the other constraints on the chiral-limit pion Bethe-Salpeter amplitude derived in Ref. from the axial-vector Ward-Takahashi identity are trivially satisfied. Using the bound state amplitudes and dressed-quark propagators we calculate the matrix elements discussed in Sec. II. Their chiral limit $`T`$-dependence is depicted in Fig. 3. As indicated by Fig. 1, the pseudoscalar and scalar bound states are massive and degenerate above $`T_c`$. Below $`T_c`$ the scalar meson residue in the scalar vertex, $`r_\sigma `$ in Eq. (37), is a little larger than the residue of the pseudoscalar meson in the pseudoscalar vertex, $`r_\pi `$ in Eq. (30). However, they are nonzero and equal above $`T_c`$, which is an algebraic consequence of $`B_00`$ and the vanishing of the subleading Dirac amplitudes. As a bona fide order parameter for chiral symmetry restoration $$f_\pi (1T/T_c)^\beta ,T/T_c<1,$$ (45) where $`\beta `$ is the zero-external-field critical-exponent for chiral symmetry restoration ($`\beta =1/2`$ in rainbow-ladder models). $`f_\pi =0`$ and $`r_\pi 0`$ for $`T>T_c`$ demonstrates that the pion disappears as a pole in the axial-vector vertex but persists as a pole in the pseudoscalar vertex. Our analysis yields $$m_\sigma (1T/T_c)^\beta ,T/T_c<1,$$ (46) within numerical errors, which is evident in Fig. 3: $`m_\sigma ^2`$ follows a linear trajectory in the vicinity of $`T_c`$. Such behaviour in the isoscalar-scalar channel might be anticipated because this channel has vacuum quantum numbers and hence the bound state is a strong interaction analogue of the electroweak Higgs boson. $`m_{\mathrm{sym}}^2`$ in Fig. 3 is the mass obtained when the chirally symmetric solution of the quark DSE is used in the BSE. ($`B_00`$ is always a solution in the chiral limit.) For $`T>T_c`$, $`m_{\mathrm{sym}}^2(T)`$ is the unique meson mass-squared trajectory. However, for $`T<T_c`$, $`m_{\mathrm{sym}}^2<0`$; i.e., the solution of the BSE in the Wigner-Weyl phase exhibits a tachyonic solution (cf. the Nambu-Goldstone phase masses: $`m_\sigma ^2>m_\pi ^2=0`$). By analogy with the $`\sigma `$-model this tachyonic mass indicates the instability of the Wigner-Weyl phase below $`T_c`$. It translates into the statement that the pressure is not maximal in this phase. Figure 4 depicts the evolution of the (common) meson mass at large $`T`$. As expected in a gas of weakly interacting quarks and gluons $$\frac{m_{\mathrm{meson}}}{2\omega _0}1^{},$$ (47) where $`\omega _0=\pi T`$ is a quark’s zeroth Matsubara frequency and “screening mass.” ### B Nonzero light current-quark masses It is straightforward to repeat the calculations of Sec. III A for nonzero current-quark masses where chiral symmetry restoration with increasing $`T`$ is exhibited as a crossover rather than a phase transition. The solutions of the inhomogeneous BSEs again exhibit a pole for all $`T`$ and we determine the bound state amplitudes from the associated homogeneous equations. Their $`T`$-dependence is characterised in Fig. 5, which shows that for $`\widehat{m}0`$ they are locally identical for $`T>\frac{4}{3}T_c`$. Figure 6 is the $`\widehat{m}0`$ analogue of Fig. 3. That the transition has become a crossover is evident in the behaviour of $`f_\pi `$. The meson masses become indistinguishable at $`T1.2T_c`$, a little before the local equivalence is manifest in Fig. 5, which is unsurprising given that the mass is an integrated quantity. The small difference between $`r_\sigma `$ and $`r_\pi `$ below $`T_c`$ is again evident and they assume a common value at the same temperature as the masses. Figure 7 illustrates the preservation of the axial-vector Ward-Takahashi identity via the mass formula of Eq. (33). The magnitude and $`T`$-dependence of both sides are equal within numerical errors above and below the crossover. $`m_R(\zeta )r_\pi ^0(\zeta )`$ is the renormalisation point independent quantity that appears in the pion’s current algebra mass formula: $`m_R(\zeta )r_\pi ^0(\zeta )`$ and $`m_R(\zeta )r_\pi (\zeta )`$ differ by $`<5`$% until $`T>0.95T_c`$. This comparison illustrates the $`T`$-domain on which the current algebra formula is valid and the analysis of Ref. . It was noted in Ref. that $$r_\pi ^0(\zeta )/f_\pi ^0(1T/T_c)^\beta ,T/T_c<1,$$ (48) which is qualitatively apparent from Figs. 3 and 7. ### C Triangle Diagrams We also studied the matrix elements describing the two-body decays discussed in Sec. II D. With every element in the calculation only known numerically this too is a challenging numerical exercise, which we simplified by a judicious choice of the external momenta. The chiral limit isoscalar-scalar-$`\pi \pi `$ coupling and width obtained from Eq. (59) are depicted in Fig. 8, which indicates that both vanish at $`T_c`$ in the chiral limit. Again this can be traced to $`B_00`$. For $`\widehat{m}0`$, the coupling reflects the crossover. However, that is moot because the width vanishes just below $`T_c`$ where the isoscalar-scalar meson mass falls below $`2m_\pi `$ and the phase space factor vanishes (see the lower panel of Fig. 6). The $`T`$-dependence of the $`\pi ^0\gamma \gamma `$ coupling, which saturates the Abelian anomaly at $`T=0`$, is calculated from Eqs. (65,68) and depicted along with the width in Fig. 9. In the chiral limit the width is identically zero because $`m_\pi =0`$ and the interesting quantity is: $`𝒯(0)=g_{\pi ^0\gamma \gamma }^0/f_\pi ^0`$. Clear in the figure is that $`𝒯(0)`$ vanishes at $`T_c`$. It vanishes with a mean field critical exponent, as is most easily inferred from Fig. 10. (An accurate calculation is possible because Eq. (44) obviates the need for a solution of the BSE.) Thus, in the chiral limit, the coupling to the dominant decay channel closes for both charged and neutral pions. These features were anticipated in Ref. . Further, as is evident in Fig. 10, our calculated $`𝒯(0)`$ is monotonically decreasing with $`T`$, supporting the perturbative O($`T^2/f_\pi ^2)`$ analysis in Ref. . For $`\widehat{m}0`$ both the coupling: $`g_{\pi ^0\gamma \gamma }/f_\pi `$, and the width exhibit the crossover with a slight enhancement in the width as $`TT_c`$ due to the increase in $`m_\pi `$. There are similarities between these results and those of Ref. although the $`T`$-dependence herein is much weaker because our pion mass approaches twice the $`T0`$ free-quark screening-mass from below, never reaching it, Eq. (47); i.e., the continuum threshold is not crossed. ## IV Summary and Conclusion We employed a renormalisation group improved rainbow-ladder truncation of the quark Dyson-Schwinger equation, and pseudoscalar and scalar Bethe-Salpeter equations to estimate the $`T`$-dependence of a range of properties that characterise correlations in these channels. The rainbow-ladder truncation is quantitatively reliable in the pseudoscalar channel. However, that is not certain in the scalar channel where the cancellations that assist in the pseudoscalar channel are not apparent. Nevertheless, we anticipate that many of the features we exposed in the scalar channel are qualitatively correct. The solutions of the inhomogeneous Bethe-Salpeter equation (BSE) exhibit poles at all $`T`$, both above and below the critical temperature, and in the chiral limit and for realistic light current-quark masses. We use the associated homogeneous BSEs to determine the masses, which correspond to the screening masses determined in simulations of lattice-QCD, and find $`m_{\sigma ,\pi }2\pi T`$ as $`T\mathrm{}`$. The BSE solutions indicate a local equivalence between the isovector-scalar and -pseudoscalar correlations above the chiral restoration transition/crossover;We refrain from asserting this of the isoscalar-scalar correlation because we cannot anticipate the effect of timelike gluon exchange contributions present in the BSE kernel in this channel. However, they are kindred to those encountered in analysing the realisation of $`U_A(1)`$ symmetry above the chiral transition. i.e., the scalar functions characterising the bound state amplitudes are identical, and from this follows equality of the masses and many of the matrix elements.The evolution to equality of the masses has been observed in numerical simulations of lattice-QCD and in the exploration of other models that accurately describe dynamical chiral symmetry breaking. Further, the axial-vector Ward-Takahashi identity and the pseudoscalar mass formula that is its corollary are valid both above and below the crossover. For realistic light current-quark masses the isoscalar-scalar meson mass does not fall below $`2m_\pi `$ until very near the transition temperature. Hence this dominant decay channel remains open almost until the phase boundary is crossed. Once it is crossed, however, only electroweak decay channels are open. In the chiral limit the anomalous two-photon coupling constant vanishes at $`T_c`$ just as does $`f_\pi `$, the coupling that determines the strength of the leptonic charged pion decay. For realistic masses, however, the widths for the leptonic charged pion mode and two-photon $`\pi ^0`$ mode remain significant in the vicinity of the crossover. The construction of a DSE-BSE truncation that allows for an improved description of the scalar channel at $`T=0`$ would provide the foundation for a significant improvement of our analysis. Employing an Ansatz for the kernel of the quark DSE whose infrared form exhibits some $`T`$-dependence, perhaps constrained by lattice simulations of the string tension, may also be interesting. However, given the results of Refs. , we do not expect such a modification to have a significant qualitative impact. In the absence of such improvements we nevertheless expect the local equivalence we have elucidated to be exhibited by all isovector chiral partners in the strong interaction spectrum. However, the explicit demonstration of this is difficult; e.g., in the $`\rho `$-$`a_1`$ complex the bound state amplitudes have eight independent amplitudes even at $`T=0`$ compared with the four in the pseudoscalar and scalar amplitudes at $`T0`$. ## Acknowledgments We acknowledge interactions with D. Blaschke and Yu.L. Kalinovsky. C.D.R. is grateful for the support and hospitality of the Special Centre for the Subatomic Structure of Matter at the University of Adelaide during a visit in which some of this work was conducted, and both C.D.R. and S.M.S. are grateful for the same from the Physics Department at the University of Rostock during a joint visit in which aspects of this work were completed. S.M.S. acknowledges financial support from the A.v. Humboldt foundation. This work was supported by the US Department of Energy, Nuclear Physics Division, under contract no. W-31-109-ENG-38, the National Science Foundation under grant nos. INT-9603385 and PHY97-22429, and benefited from the resources of the National Energy Research Scientific Computing Center. ## Collected Formulae ### 1 Scalar Vertex The inhomogeneous ladder-like Bethe-Salpeter equation for the zeroth Matsubara mode of the $`0^{++}`$ vertex is $`\mathrm{\Gamma }_\mathrm{s}^\alpha (p_{\omega _k};P_0;\zeta )=Z_4{\displaystyle \frac{1}{2}}\tau ^\alpha \mathrm{𝟏}`$ (49) $`{\displaystyle _{l,q}^{\overline{\mathrm{\Lambda }}}}{\displaystyle \frac{4}{3}}g^2D_{\mu \nu }(p_{\omega _k}q_{\omega _l})`$ (50) $`\times `$ $`\gamma _\mu S(q_{\omega _l}^+)\mathrm{\Gamma }_\mathrm{s}^i(q_{\omega _l};P_{\mathrm{\Omega }_n};\zeta )S(q_{\omega _l}^{})\gamma _\nu ,`$ (51) where $`\alpha =0,1,2,3`$ with $`\tau ^0=\mathrm{diag}(1,1)`$. (NB: In this truncation the isoscalar and isovector states are degenerate, which exemplifies our observation that the ladder-like truncation is accurate for vector mesons but requires improvement before it is quantitatively reliable in the $`0^{++}`$ sector.) The solution has the form $`\mathrm{\Gamma }_\mathrm{s}^i(p_{\omega _k};\stackrel{}{P})={\displaystyle \frac{1}{2}}\tau ^\alpha \mathrm{𝟏}[E_\mathrm{s}(p_{\omega _k};\stackrel{}{P})+i\stackrel{}{\gamma }\stackrel{}{p}G_\mathrm{s}^{}(p_{\omega _k};\stackrel{}{P})`$ (52) $`+`$ $`i\gamma _4\omega _kG_\mathrm{s}^{}(p_{\omega _k};\stackrel{}{P})+i\stackrel{}{\gamma }\stackrel{}{P}\stackrel{}{p}\stackrel{}{P}F_\mathrm{s}(p_{\omega _k};\stackrel{}{P})].`$ (53) (NB: Here the requirement that the neutral mesons be charge conjugation eigenstates shifts the $`\stackrel{}{p}\stackrel{}{P}`$ term cf. the $`0^+`$ amplitude.) The scalar functions in Eq.(51) exhibit a simple pole at $`\stackrel{}{P}^2+m_\sigma ^2=0`$: $$\mathrm{\Gamma }_\mathrm{s}^i(p_{\omega _k};\stackrel{}{P})=\frac{r_\sigma (\zeta )}{\stackrel{}{P}^2+m_\sigma ^2}\mathrm{\Gamma }_\sigma ^i(p_{\omega _k};\stackrel{}{P})+\mathrm{regular},$$ (54) where $`\mathrm{\Gamma }_\sigma ^i(p_{\omega _k};\stackrel{}{P})`$ is the canonically normalised, $`0^{++}`$ bound state Bethe-Salpeter amplitude: $`2\delta ^{\alpha \beta }\stackrel{}{P}=\mathrm{tr}{\displaystyle _{l,q}^{\overline{\mathrm{\Lambda }}}}\{\mathrm{\Gamma }_\sigma ^\alpha (q_{\omega _l};\stackrel{}{P}){\displaystyle \frac{S(q_{\omega _l}^+)}{\stackrel{}{P}}}\mathrm{\Gamma }_\sigma ^\beta (q_{\omega _l};\stackrel{}{P})S(q_{\omega _l}^{})`$ (56) $`+\mathrm{\Gamma }_\sigma ^\alpha (q_{\omega _l};\stackrel{}{P})S(q_{\omega _l}^+)\mathrm{\Gamma }_\sigma ^\beta (q_{\omega _l};\stackrel{}{P}){\displaystyle \frac{S(q_{\omega _l}^{})}{\stackrel{}{P}}}\left\}\right|_{\stackrel{}{P}^2=m_\mathrm{s}^2}.`$ The dressed-quark propagator and canonically normalised Bethe-Salpeter amplitudes make possible the definition of the impulse approximation to the isoscalar-scalar-$`\pi \pi `$ matrix element: $`\stackrel{}{p_1}^2=m_\pi ^2=\stackrel{}{p_2}^2`$, $`(\stackrel{}{p}=\stackrel{}{p_1}+\stackrel{}{p_2})^2=m_\sigma ^2`$, $`g_{\sigma \pi \pi }:=\pi (\stackrel{}{p_1})\pi (\stackrel{}{p_2})|\sigma (\stackrel{}{p})=`$ (59) $`2N_c\mathrm{tr}_D{\displaystyle _{l,q}^{\overline{\mathrm{\Lambda }}}}\mathrm{\Gamma }_\sigma (k_{\omega _l};\stackrel{}{p})S_u(k_{++})`$ $`\times i\mathrm{\Gamma }_\pi (k_{0+};\stackrel{}{p_1})S_u(k_+)i\mathrm{\Gamma }_\pi (k_0;\stackrel{}{p_2})S_u(k_{}),`$ $`k_{\alpha \beta }=k_{\omega _l}+(\alpha /2)\stackrel{}{p_1}+(\beta /2)\stackrel{}{p_2}`$, with only the trace over Dirac indices remaining. ### 2 Neutral Pion Decay An efficacious Ansatz for the dressed-quark-photon coupling at $`T=0`$ is: $`i\mathrm{\Gamma }_\mu (\mathrm{}_1,\mathrm{}_2)=i\mathrm{\Sigma }_A(\mathrm{}_1^2,\mathrm{}_2^2)\gamma _\mu `$ (63) $`+(\mathrm{}_1+\mathrm{}_2)_\mu \left[\frac{1}{2}i\gamma (\mathrm{}_1+\mathrm{}_2)\mathrm{\Delta }_A(\mathrm{}_1^2,\mathrm{}_2^2)+\mathrm{\Delta }_B(\mathrm{}_1^2,\mathrm{}_2^2)\right];`$ $`\mathrm{\Sigma }_F(\mathrm{}_1^2,\mathrm{}_2^2)=\frac{1}{2}[F(\mathrm{}_1^2)+F(\mathrm{}_2^2)],`$ $`\mathrm{\Delta }_F(\mathrm{}_1^2,\mathrm{}_2^2)={\displaystyle \frac{F(\mathrm{}_1^2)F(\mathrm{}_2^2)}{\mathrm{}_1^2\mathrm{}_2^2}},`$ where $`F=A,B`$; i.e., the scalar functions in the dressed-quark propagator: $`S^1(p)=i\gamma pA(p^2)+B(p^2)`$, so that this model is completely determined by $`S(p)`$. Improvements of this Ansatz, such as those canvassed in Ref. , do not have a qualitatively significant effect in the present context. The dressed-quark-photon vertex is a necessary element in the calculation of the impulse approximation to the $`\pi ^0\gamma \gamma `$ amplitude. At $`T=0`$ the anomalous contribution to the divergence of the axial-vector vertex is saturated by the pseudoscalar piece of the pion Bethe-Salpeter amplitude $`\widehat{T}_{\mu \nu }(k_1,k_2)=\mathrm{tr}{\displaystyle _{l,q}^{\overline{\mathrm{\Lambda }}}}S(q_1)\gamma _5\tau ^3iE_\pi (\widehat{q};P)`$ (64) $`\times `$ $`S(q_2)i𝒬\mathrm{\Gamma }_\mu (q_2,q_{12})S(q_{12})i𝒬\mathrm{\Gamma }_\nu (q_{12},q_1),`$ (65) where $`𝒬=\mathrm{diag}(2/3,1/3)`$ and herein $`k_1=(\stackrel{}{k}_1,0)`$, $`k_2=(\stackrel{}{k}_2,0)`$, $`P=k_1+k_2`$, $`q_1=q_{\omega _l}k_1`$, $`q_2=q_{\omega _l}+k_2`$, $`\widehat{q}=\frac{1}{2}(q_1+q_2)`$, $`q_{12}=q_{\omega _l}k_1+k_2`$. Using Eq. (63) to evaluate Eq. (65) for real photons at $`T=0`$ one obtains $$\widehat{T}_{\mu \nu }(k_1,k_2)=\frac{\alpha _{\mathrm{em}}}{\pi }ϵ_{\mu \nu \rho \sigma }k_{1\rho }k_{2\sigma }𝒯(0),$$ (66) with in the chiral limit $$f_\pi ^0𝒯(0):=g_{\pi ^0\gamma \gamma }=1/2.$$ (67) At nonzero $`T`$ the tensor structure of Eq. (66) survives to the extent that, with our choice of $`k_1`$, $`k_2`$, it ensures one of the photons is longitudinal (a plasmon) and the other transverse. In this case we determine the $`T`$-dependence using $$\widehat{T}_{i4}(k_1,k_2)=\frac{\alpha _{\mathrm{em}}}{\pi }(\stackrel{}{k_1}\times \stackrel{}{k_2})_i𝒯(0)$$ (68) and a generalisation of Eq. (63) to nonzero $`T`$: $`i\stackrel{}{\mathrm{\Gamma }}(q_{\omega _{l_1}},q_{\omega _{l_2}})=\mathrm{\Sigma }_A(q_{\omega _{l_1}}^2,q_{\omega _{l_2}}^2)i\stackrel{}{\gamma }`$ (69) $`+`$ $`(\stackrel{}{q}_1+\stackrel{}{q}_2)[{\displaystyle \frac{1}{2}}iG(q_{\omega _{l_1}},q_{\omega _{l_2}})+\mathrm{\Delta }_B(q_{\omega _{l_1}}^2,q_{\omega _{l_2}}^2)],`$ (70) $`i\mathrm{\Gamma }_4(q_{\omega _{l_1}},q_{\omega _{l_2}})=\mathrm{\Sigma }_C(q_{\omega _{l_1}}^2,q_{\omega _{l_2}}^2)i\gamma _4`$ (72) $`+(\omega _{l_1}+\omega _{l_2})[{\displaystyle \frac{1}{2}}iG(q_{\omega _{l_1}},q_{\omega _{l_2}})+\mathrm{\Delta }_B(q_{\omega _{l_1}}^2,q_{\omega _{l_2}}^2)],`$ $`G(q_{\omega _{l_1}},q_{\omega _{l_2}})=\stackrel{}{\gamma }(\stackrel{}{q}_1+\stackrel{}{q}_2)\mathrm{\Delta }_A(q_{\omega _{l_1}}^2,q_{\omega _{l_2}}^2)`$ (74) $`+\gamma _4(\omega _{l_1}+\omega _{l_2})\mathrm{\Delta }_C(q_{\omega _{l_1}}^2,q_{\omega _{l_2}}^2),`$ which satisfies the vector Ward-Takahashi identity $$(q_{\omega _{l_1}}q_{\omega _{l_2}})_\mu i\mathrm{\Gamma }_\mu (q_{\omega _{l_1}},q_{\omega _{l_2}})=S^1(q_{\omega _{l_1}})S^1(q_{\omega _{l_2}}).$$ (75)
warning/0001/gr-qc0001101.html
ar5iv
text
# General Ether Theory ## 1 Introduction The purpose of the present work is to present an alternative metric theory of gravity. The Lagrangian of the theory $$L=L_{GR}(8\pi G)^1(\mathrm{{\rm Y}}g^{00}\mathrm{\Xi }(g^{11}+g^{22}+g^{33}))\sqrt{g}$$ is very close to the GR Lagrangian, and in the limit $`\mathrm{\Xi },\mathrm{{\rm Y}}0`$ we obtain the classical Einstein equations. The key point is that this Lagrangian may be derived starting with a few assumptions about the “ether” – a classical medium in a classical Newtonian background with Euclidean space and absolute time $`^3`$. We need only a few general principles: a Lagrange formalism and its relation with standard conservation laws. The gravitational field $`g^{\mu \nu }`$ is defined by the “general” steps of freedom of the ether – density $`\rho `$, velocity $`v^i`$, pressure $`p^{ij}`$.<sup>1</sup><sup>1</sup>1As usual, we use latin indices for three-dimensional indices and greek indices for four-dimensional indices. We also use the notation $`\widehat{g}^{\mu \nu }=g^{\mu \nu }\sqrt{g}`$. The matter fields describe its material properties. What explains the Einstein equivalence principle is that the ether is universal: all fields describe properties of the ether, there is no external matter. Therefore, observers are also only excitations of the ether, unable to observe some of the ether properties. This explains that we are unable to observe all steps of freedom of the ether. We need no artificial conspiracy or highly sophisticated model to obtain relativistic symmetry in an ether theory. The only difference to GR are two additional terms which depend on the preferred coordinates $`X^i,T`$. In “weak” covariant formulation (with the preferred coordinates handled as “fields” $`X^i(x),T(x)`$) we obtain: $$L=L_{GR}(8\pi G)^1(\mathrm{{\rm Y}}g^{\mu \nu }T_{,\mu }T_{,\nu }\mathrm{\Xi }g^{\mu \nu }\delta _{ij}X_{,\mu }^iX_{,\nu }^j)\sqrt{g}$$ Instead of no equation for the preferred coordinates, we obtain a well-defined general equation for these coordinates: the classical conservation laws, which appear to be the harmonic coordinate condition. But this changes a lot. It is, essentially, a paradigm shift as described by Kuhn . We revive the metaphysics of Lorentz ether theory in its full beauty. This requires the reconsideration of the whole progress made in fundamental physics in this century. Therefore, after derivation of the theory and their comparison with existing theories of gravity we reconsider different domains of science from point of view of the new paradigm. Of course, this may be only a raw overview, a program for future research instead of a summary of results. But this raw overview does not suggest serious problems for the new paradigm, while essential problems of the relativistic paradigm disappear. The preferred background leads to well-defined local energy and momentum conservation laws. Moreover, the additional terms seem to be useful to solve cosmological problems. The $`\mathrm{\Xi }`$-term defines a nice homogeneous dark matter candidate. The $`\mathrm{{\rm Y}}`$-term is even more interesting: it avoids the big bang singularity and leads instead to a bounce. Such a bounce makes the cosmological horizon much larger and therefore solves the cosmological horizon problem without inflation. This term stops also the gravitational collapse immediately before horizon formation. Because of the underlying Euclidean symmetry, the flat universe is the only homogeneous universe. Therefore, GET is not only in agreement with observation, but allows to solve some serious cosmological problems solved today by inflation theory. A very strong argument in favour of GET is the violation of Bell’s inequality. In our opinion, it is a very simple and decisive proof of the existence of a preferred foliation – as simple and decisive as possible in fundamental physics. Only if we try to avoid this simple conclusion, the issue becomes complicate – we have to reject simple fundamental principles like the EPR criterion of reality or causality. In our opinion, there is a lot of confusion in this question. For example, it is often assumed that the EPR criterion is in contradiction with quantum theory. But the existence of Bohmian mechanics proves that there is no such contradiction. In quantum field theory the reintroduction of a preferred frame does not lead to problems. Instead, it allows to generalize Bohmian mechanics into the relativistic domain and clarifies the choice of the Fock space in semiclassical field theory. The “ether hypothesis” suggests also a simple solution for the problem of non-renormalizability. Technically, this solution is already known as “effective field theory”. In this concept, it is assumed that below a certain cutoff scale the theory becomes really different. This concept has two features where GET suggests modification: First, in the standard concept the nature of the theory below this cutoff remains completely unspecified. Instead, GET suggests a well-defined framework: canonical quantum theory, Newtonian space-time, and some “atomic ether theory”. Second, the cutoff length is supposed to be the Planck length. Instead, GET makes a prediction which is inconsistent with Planck length – the cutoff is defined by $`g^{00}\sqrt{g}V_{cutoff}=1`$. As a consequence, the cutoff length seems to increase in a homogeneous “expanding” universe. Remarkably, all these results are only side-effects. It was not the original intention of the author to revive old ether theory. Instead, the author shares the common admiration for the beauty of GR. It was also not the intention to solve cosmological problems – they have been considered only after the derivation of the Lagrangian. We also have not tried to save the EPR criterion or to generalize Bohmian mechanics. The original motivation was different. It was a quantum gravity thought experiment which has convinced the author that a Newtonian framework is necessary. The question is if a “one world theory” as GR is sufficient to describe superpositions of different gravitational fields, or if such a superposition depends on relations between the superposed fields, their “relative position”. To decide this question, we consider a simple interaction of a superposition of quasiclassical fields with a test particle. We observe a transition probability which depends on such relative information. To describe such transition probabilities in quantum gravity it seems necessary to introduce a common background. Last not least, it seems necessary to criticize some aspects of the relativistic paradigm for quantum gravity. “Because of the lack of data, quantum gravity is strongly influenced by philosophical prejudices of the researchers” , therefore, these prejudices have to be considered. We use Rovelli as a base for this consideration. It includes an excellent methodological part we agree with. We criticize his relativistic argumentation and argue that our consideration is in much better agreement with the proposed methodology. ## 2 General Ether Theory Let’s now define general ether theory. We have a Newtonian framework – absolute Euclidean space with orthonormal coordinates $`X^i`$ and absolute time T. We have also classical causality – causal influence $`AB`$ between events A and B is possible only if $`T(A)T(B)`$. The ether is described by steps of freedom which are usual in condensed matter theory: there is an “ether density” $`\rho (X,T)`$, an “ether velocity” $`v^i(X,T)`$ and an “ether pressure” $`p^{ij}(X,T)`$. As usual for a density, $`\rho >0`$. These steps of freedom define the gravitational field. The theory is a metric theory of gravity, and the metric $`g_{\mu \nu }`$ is defined algebraically by the following formula: $`g^{00}\sqrt{g}`$ $`=`$ $`\rho `$ $`g^{i0}\sqrt{g}`$ $`=`$ $`\rho v^i`$ $`g^{ij}\sqrt{g}`$ $`=`$ $`\rho v^iv^j+p^{ij}`$ This formula is a variant of the ADM decomposition. Especially, $`v^i`$ is the ADM shift vector. Because the density $`\rho `$ is always positive, this formula defines a Lorentz metric if and only if the tensor $`p^{ij}`$ is negative definite. Therefore, we make the additional assumption that $`p^{ij}`$ is negative definite. ### 2.1 The material properties of the ether These are not all steps of freedom of the ether. Instead, there are other steps of freedom, the “material properties” $`\phi ^m(X,T)`$ of the ether. But these steps of freedom are not defined by GET. Instead, GET is a general theory only, it describes only a few general properties, not all properties of the ether. It is, therefore, a meta-theory, a general scheme for an ether theory. Different ether theories can fit into this scheme. Ether theories with well-defined material properties and material laws we name “complete ether models”. This is in no way strange for a theory of gravity. All metric theories of gravity are general schemes in the same sense. They do not specify the matter steps of freedom and the matter Lagrangian. Nonetheless, they specify an essential and very important property of the matter Lagrangian – that it fulfils the Einstein equivalence principle. Thus, the meta-theoretical character is a common feature of theories of gravity. GET in some sense explains this subdivision into a universal gravitational field and matter fields. Indeed, there is a similar subdivision in condensed matter theory – the subdivision between the few basic steps of freedom (like density, velocity, pressure) which are common for very different materials and the “material properties” which differ for different materials. In GET this subdivision fits with the subdivision into gravity and matter fields: Density, velocity and pressure are used to describe the gravitational field, while the other material properties describe the matter fields. This is an essential difference to classical ether theory. In the classical concept, the ether is assumed to be something different from usual matter. In GET, usual matter is described by continuous fields, and these fields describe various properties of the ether. But we not only assume that there are material properties of the ether described by some matter fields. We assume more: all matter fields describe material properties of the ether. There is no ethr-external matter: ###### Axiom 1 (universality) There is nothing except the ether. All fields describe steps of freedom of the ether. Thus, the complete ether model is the theory of everything. ### 2.2 Conservation laws In our covariant formalism, the conservation laws are the Euler-Lagrange equations for the preferred coordinates $`X^\mu `$ (see (A.2)). Now, let’s try to find these conservation laws. The main hypothesis is that these conservation laws coinside with the classical conservation laws we know from condensed matter theory. ###### Axiom 2 (continuity equation) The mass of the ether is conserved. This conservation is described by the classical continuity equation: $$_t\rho +_i(\rho v^i)=0$$ (1) The other important equation of classical condensed matter theory is the Euler equation. It is the conservation law for momentum: ###### Axiom 3 (Euler equation) The momentum of the ether is conserved. This conservation is described by the classical Euler equation: $$_t(\rho v^j)+_i(\rho v^iv^j+p^{ij})=0$$ (2) Note that we have no terms for external forces or interaction with external matter. The reason is that we have already incorporated here the universality axiom – there is no momentum exchange with external matter, there are no external forces. All “matter fields” are “material properties” of the ether. These are already all ether equations specified by GET. All other equations are “material laws” of the ether, they depend on the “material properties” $`\phi ^m`$ which have to be defined only by the complete ether model. GET does not specify them. Now, a key observation is what happens if if we rewrite the classical conservation laws as equations for the effective metric $`g_{\mu \nu }`$. We obtain a well-known equation – the harmonic condition: $$\mathrm{}X^\nu =_\mu (g^{\mu \nu }\sqrt{g})=0$$ ### 2.3 Lagrange formalism A major assumption is that we have a Lagrange formalism. We use the covariant formulation of the theory: the preferred coordinates $`X^\mu `$ are considered as fields, the Lagrangian depends on the fields $`X^\mu (x)`$ in a covariant way (see §A.1). In this formalism, the conservation laws are the Euler-Lagrange equations for the preferred coordinates (see (A.2)). On the other hand, we have already found the conservation laws and observed that they may be written as equations for the preferred coordinates. Thus, it seems reasonable to assume that they are proportional: $$\frac{\delta S}{\delta X^\mu }=\gamma _\mu \mathrm{}X^\mu $$ Now, the coefficients $`\gamma _\mu `$ may be different. We can use Euclidean symmetry to argue that $`\gamma _1=\gamma _2=\gamma _3`$, but there is no reason to suppose a relation between $`\gamma _0`$ and the $`\gamma _i`$. Instead of the $`\gamma _\mu `$ we introduce introduce a diagonal matrix $`\gamma _{\mu \nu }`$ with $`\gamma _{\mu \mu }=4\pi G\gamma _\mu `$. The factor $`4\pi G`$ is well-known from GR, and it seems natural to introduce it here: in this case, the two constants $`\mathrm{{\rm Y}}=\gamma _{00}`$, $`\mathrm{\Xi }=\gamma _{ii}`$ appear in a similar way as Einstein’s cosmological constant $`\mathrm{\Lambda }`$. Now we can formulate the ###### Axiom 4 (Lagrange formalism) There exists a “weak covariant” Lagrange formalism so that the Euler-Lagrange equations for the preferred coordinates $`X^\mu `$ and the classical conservation laws for the ether $`\mathrm{}X^\mu =0`$ are related in the following way: $$\frac{\delta S}{\delta X^\mu }=(4\pi G)^1\gamma _{\mu \nu }\mathrm{}X^\nu $$ (3) Now, let’s find the general Lagrangian which fulfils this assumption. ###### Theorem 1 The general Lagrangian for GET is $$L=(8\pi G)^1\gamma _{\mu \nu }g^{\mu \nu }\sqrt{g}+L_{GR}(g_{\mu \nu })+L_{matter}(g_{\mu \nu },\phi ^m)$$ (4) Proof: First, we find a Lagrangian which fulfils the condition (3): $$L_{GET}=(8\pi G)^1\gamma _{\mu \nu }X_{,\alpha }^\mu X_{,\beta }^\nu g^{\alpha \beta }\sqrt{g}$$ For the difference we obtain $$\frac{\delta (LL_{GET})}{\delta X^\mu }=0$$ Thus, the remaining part is “strong” covariant, that means, it is not only covariant, but does not depend on the preferred coordinates $`X^\mu `$ too. But this is the classical requirement for the Lagrangian of general relativity. Thus, we can identify the difference with the classical Lagrangian of general relativity. $$L=L_{GET}(g_{\mu \nu },X^\mu )+L_{GR}(g_{\mu \nu })+L_{matter}(g_{\mu \nu },\phi ^m)$$ In the preferred coordinates $`L_{GET}`$ may be rewritten as $$L_{GET}=(8\pi G)^1(\mathrm{{\rm Y}}g^{00}\mathrm{\Xi }(g^{11}+g^{22}+g^{33}))\sqrt{g}$$ or in a more compact form as $$L_{GET}=(8\pi G)^1\gamma _{\mu \nu }g^{\mu \nu }\sqrt{g}$$ This proves the theorem. It should be noted that we have no theoretical reason to fix the signs for the cosmological constants $`\mathrm{\Xi },\mathrm{{\rm Y}},\mathrm{\Lambda }`$. Their values should be fixed by observation. ## 3 Simple properties Now we have defined GET and can describe its properties. First, let’s write down the other Euler-Lagrange equations. As equations for $`g^{\mu \nu }`$ we obtain the Einstein equations with two additional non-covariant terms: $$G_\nu ^\mu =8\pi G(T_m)_\nu ^\mu +(\mathrm{\Lambda }+\gamma _{\kappa \lambda }g^{\kappa \lambda })\delta _\nu ^\mu 2g^{\mu \kappa }\gamma _{\kappa \nu }$$ (5) As in GR, the equations for the matter fields $`\phi ^m`$ depend on the matter Lagrangian and remain unspecified. The expression of the GET Lagrangian in terms of the original ether variables is quite nice: $$4\pi G\mathrm{\Xi }^1L_{GET}=\frac{1}{2}(\rho |v|^2+p^{ii}\mathrm{{\rm Y}}\mathrm{\Xi }^1\rho )$$ ### 3.1 Energy-momentum tensor Now, we have derived the GET Lagrangian using assumptions about the conservation laws. Therefore, to write dowm the energy-momentum tensor is easy: $$T_\nu ^\mu =(4\pi G)^1\gamma _{\nu \kappa }g^{\kappa \mu }\sqrt{g}$$ In this form, the tensor does not depend on the material properties $`\phi ^m`$ of the ether. But how is this energy-momentum tensor related with the usual energy-momentum tensor for the matter fields? Now, the answer is simple. We have to multiply the GET variant of the Einstein equation (5) with $`\sqrt{g}`$ and obtain the following decomposition of the full energy-momentum tensor: $$T_\nu ^\mu =(T_m)_\nu ^\mu \sqrt{g}+(8\pi G)^1\left((\mathrm{\Lambda }+\gamma _{\kappa \lambda }g^{\kappa \lambda })\delta _\nu ^\mu G_\nu ^\mu \right)\sqrt{g}$$ Thus, we obtain immediately what is missed in GR: an energy-momentum tensor for the gravitational field. ### 3.2 Constraints If we want to formulate an initial value problem for GR as well as GET, we cannot simply define the initial values $`g_{\mu \nu }^0(x)=g_{\mu \nu }(x,0)`$ and $`k_{\mu \nu }^0(x)=_tg_{\mu \nu }(x,t)|_{t=0}`$. Instead, we obtain the problem that the four equations $$G_\mu ^0=\mathrm{}$$ do not contain second order derivatives in time, that means, they define constraints for the initial values. This is a common property in above theories, because the additional terms of GET do not add second order derivatives of the $`g_{\mu \nu }`$. Moreover, in GET the four conservation laws are also only first order in time. Nonetheless, their character is completely different. In GR, these constraints play a very special role in the ADM Hamilton formalism – the energy H itself is a constraint. This is a consequence of the covariance of the GR Lagrangian. In GET, as we have already seen, we have a well-defined local energy and momentum density. Even if the constraints are much more harmless in GET, they remain to be constraints, which is not very nice. But there is some interesting insight into the nature of the constraints in GET which has been found for GR in harmonic coordinates by Choquet-Bruhat . This insight was important for his proof of local existence and uniqueness theorems for GR. It is remarkable in itself that this proof has been done in harmonic coordinates. First, as has been observed by Lanczos , the Ricci tensor essentially simplifies in harmonic coordinates: $$R_{\mu \nu }^{(h)}=\frac{1}{2}g^{\alpha \beta }\frac{^2g_{\mu \nu }}{x^\alpha x^\beta }+H_{\mu \nu }$$ where $`H_{\mu \nu }`$ does not contain second derivatives of the metric. Now, in GET we have the harmonic condition $`\mathrm{\Gamma }^\mu =_\nu (g^{\mu \nu }\sqrt{g})=0`$ as an equation. Therefore, for the initial values we have $$\mathrm{\Gamma }^\mu (0,x)=0_t\mathrm{\Gamma }^\mu (0,x)=0$$ The second condition contains a second order time derivative. But this is true also for the Ricci tensor in harmonic coordinates. Now, if we use the appropriate combination of this second initial condition and the equation in harmonic coordinates, we obtain the four other first order constraints (see , Lemma 22). Thus, all constraints are closely related to the harmonic condition. The ether interpretation gives additional insight. The point is that in this interpretation the components $`\rho v^i=g^{0i}\sqrt{g}`$ are already velocities. A second order equation for these components would be a third order equation for the ether particles them-self. Therefore, it is very natural that there are no such third order terms in the equations them-self. ## 4 Derivation of the Einstein equivalence principle The Einstein equivalence principle (EEP) is an immediate consequence of the GET Lagrange density: the matter Lagrangian is covariant in the strong sense, does not depend on the preferred coordinates $`X^\mu `$. The question we want to consider here is if there are generalizations of the GET axioms so that the EEP remains correct. At first we have to formulate the EEP in an appropriate way. The equations for matter do not depend on the preferred coordinates. In our covariant formalism there is a natural way to do formulate this property. The “equations for the matter” we identify with the Euler-Lagrange equations for $`\phi ^m`$, and the property that they do not depend on the preferred coordinates $`X^\mu `$ can be written as $$\frac{\delta }{\delta X^\mu }\frac{\delta S}{\delta \phi ^m}=0$$ To obtain a proof, let’s look how this property may be proven for the GET Lagrangian: ###### Theorem 2 (Einstein equivalence principle) Let L be a weak covariant Lagrangian with the conservation laws $$_\mu T_\nu ^\mu =0.$$ If the conservation laws do not depend on the variables $`\phi ^m`$, then the Einstein equivalence principle holds for these variables. Proof: The conservation laws in the weak covariant formalism are defined as $$\frac{\delta S}{\delta X^\mu }=_\nu T_\mu ^\nu =0$$ The EEP follows immediately: $$\frac{\delta }{\delta X^\mu }\frac{\delta S}{\delta \phi ^m}=\frac{\delta }{\delta \phi ^m}\frac{\delta S}{\delta X^\mu }=\frac{\delta }{\delta \phi ^m}_\nu T_\mu ^\nu =0$$ As we see, the first property we need is that the material properties are not used in the conservation laws. This property depends on our choice of $`p^{ij}`$ as an independent variable. But this is only a technical question. More important is the universality axiom – that the matter fields $`\phi ^m`$ describe only “material properties” of the ether. External steps of freedom, that means other, non-ether fields which interact with the ether have some momentum exchange with the ether. Therefore, there will be some interaction terms in the momentum conservation laws. But this may be partially weakened. If some steps of freedom $`\psi ^n`$ are external forces or external matter, while other steps of freedom $`\phi ^m`$ describe material properties of the ether, than the EEP does not hold for the external steps of freedom, but remains valid for the material properties. The proof remains the same. The other property is that the conservation laws are the Euler-Lagrange equations for the preferred coordinates: $$\frac{\delta S}{\delta X^\mu }=_\nu T_\mu ^\nu $$ Now, this may be generalized for the case where we have explicit dependencies on the preferred coordinates, and, therefore, no conservation laws. All we need is that the equation does not depend on the material properties $`\phi ^m`$. Therefore, we conclude that the EEP holds for material properties $`\phi ^m`$ even in more general situations, if we have other external fields, external forces, even explicit dependencies of the Lagrangian from the coordinates. Let’s summarize: The EEP holds for a step of freedom $`\phi ^m`$ only if it describes a material property of the ether. The universality axiom explains why the EEP holds for all matter fields. ### 4.1 Higher order approximations in a Lagrange formalism Let’s consider now another possibility for generalization. We consider the situation where we have to consider different approximations for a Lagrange formalism. All we assume is that all approximations are consistent. For a condensed matter theory that means that the conservation laws are valid. Let’s compare now two approximations. For the approximation $`S=S_0+S_1`$ we have $$\frac{\delta S_0}{\delta X^\mu }=_\nu T_\mu ^\nu $$ as well as $$\frac{\delta S}{\delta X^\mu }=_\nu T_\mu ^\nu $$ Now, in this situation we do not even need that these conservation laws do not depend on some fields $`\phi ^m`$. All we need now is that the two approximations of the conservation laws are identical. This leads immediately to the Einstein equivalence principle for the additional part of the Lagrangian: $$\frac{\delta S_1}{\delta X^\mu }=0$$ This consideration suggests that it may be even easier to detect relativistic symmetry in the higher order approximations. Sequences $`S_n`$ of approximations appear in effective field theory. ### 4.2 Weakening the assumptions about the Lagrange formalism We have obtained the conservation laws with reference to condensed matter theory. Then we have identified them with the conservation laws from the weak covariant formalism. This is a quite natural, but non-trivial identification. There are many different variants of the conservation laws, and even if they are equivalent if the equations of motion are fulfilled, their functional dependencies differ. Therefore, it would be nice to weaken this assumption. Unfortunately, it seems impossible to prove something without an explicit assumption which relates a certain conservation law with the Euler-Lagrange equations. The property “there exists a Lagrange formalism” for some equivalent system of equations is too weak. The problem is that there are various methods of transformation of a given system of equations – multiplying them with “integrating factors”, Lagrange multipliers, replacement of fields by potentials. That’s why a general method which allows to decide if a given set of equations is equivalent to a system of Euler-Lagrange equations is not known . That means, we are not even able to find all Lagrange formalisms for a given set of equations. It seems natural to assume that the equations already have the form of Euler-Lagrange equations. Such systems of equations are “self-adjoint” and have been considered in detail . Especially there exist standard methods to construct Lagrange densities which are also tests if the system is self-adjoint . Thus, let’s assume that the conservation laws are part of such a self-adjoint system of equations, that means, are Euler-Lagrange equations for some variables $`c^\mu `$: $$\frac{\delta S}{\delta c^\mu }=_\nu T_\mu ^\nu $$ This allows to derive a similar symmetry property of the equations for $`\phi ^m`$: $$\frac{\delta }{\delta c^\mu }\frac{\delta S}{\delta \phi ^m}=\frac{\delta }{\delta \phi ^m}\frac{\delta S}{\delta c^\mu }=\frac{\delta }{\delta \phi ^m}_\nu T_\mu ^\nu =0$$ Thus, we have a symmetry group with four continuous parameters, only the relation between these parameters and the coordinates has been lost. ### 4.3 Explanatory power of the derivation Last not least, it should be noted that the derivation of the EEP has very high explanatory power. First, as we have seen, it is based on a few very general principles. We do not need any special assumptions about the ether, no “mechanical explanation”, no strange “mechanism”, no “conspiracy”. We make non-trivial assumptions, but these non-trivial assumptions are very natural for a condensed matter theory. Evidence for the high explanatory power is that we can describe the proof in a simple verbal way: The conservation laws can be understood as equations for the preferred coordinates. Now, the conservation laws do not depend on the material properties. Therefore, because of the principle “action equals reaction”, the equations for the material properties do not depend on the preferred coordinates. ## 5 Does usual matter fit into the GET scheme? If we compare the Einstein equations with usual hydro-dynamical equations, they look very different. The Einstein equations depend on second order derivatives of the $`g_{\mu \nu }`$. At a first look, there seems to be no chance to unify them. GET suggests such a way. Indeed, in the derivation of GET we have used only a few general properties of the ether. None of these assumptions is obviously wrong for usual condensed matter theory. The material properties of the GET ether remain unspecified. We cannot even tell if the ether is solid or liquid.<sup>2</sup><sup>2</sup>2It may be assumed that the condition that the pressure tensor $`p^{ij}`$ is negative definite tells something about the ether – that the ether is a material with negative pressure. But the only reason for naming the tensor field $`p^{ij}`$ “pressure” is that it appears like usual pressure appears in the usual Euler equation. And this allows to identify $`p^{ij}`$ with usual pressure only modulo a constant. This suggests that usual condensed matter may be described by a GET-like Lagrangian. ### 5.1 The role of the Einstein Lagrangian The problem with the second derivatives which appear in the Einstein equations can be easily solved: we have identified the “remaining part” of the GET Lagrangian with the Einstein Lagrangian because of its strong covariance. None of the GET axioms requires to include the Einstein-Hilbert term $`L_{GR}=R\sqrt{g}`$ into the GET Lagrangian. It is simply a possible term in the GET Lagrangian, not a necessary one. The same holds for covariant terms with higher order.<sup>3</sup><sup>3</sup>3As described by Weinberg “there’s no reason in the world to suppose that the Lagrangian does not contain all the higher terms with more factors of the curvature and/or more derivatives, all of which are suppressed”. Moreover, in comparison with the three “cosmological terms” $`g^{00}\sqrt{g},g^{ii}\sqrt{g},\sqrt{g}`$ of GET the Einstein-Hilbert term is a higher order term. Therefore, in the first approximation for usual condensed matter the Einstein-Hilbert term may be simply omitted. In this case, the GET equation becomes an algebraic relation between the gravitational field defined by $`\rho ,v^i,p^{ij}`$ and the material properties $`\phi ^m`$. This is already much more close to usual condensed matter theory. On the other hand, GET suggests that the Einstein-Hilbert Lagrangian should be used in higher order approximations of condensed matter theory. The physical meaning of curvature-dependent term in condensed matter theory is easy to understand: if curvature is zero, then there exists an undistorted reference state which remains unchanged in time. Therefore, curvature describes inner stress and its change in time. ### 5.2 The role of Lorentz symmetry Another property seems to be much more in contradiction with usual hydrodynamics – the Lorentz invariance of the GET Lagrangian. The physical meaning of this Lorentz invariance is not clear. One possibility is that it has none, and it simply as a consequence of the fact that there are not very much possibilities for quadratic Lagrangians, therefore, symmetries appear more or less by accident. With another choice of the constants, especially $`Xi<0`$, we would obtain an SO(4) symmetry, also without any physical meaning. Nonetheless, a consequence is that we cannot describe with GET a Galilean-invariant theory. On the other hand, we can use GET to describe special-relativistic condensed matter. The reverse method would be to describe a Galilean invariant theory as the limit $`\mathrm{\Xi }0`$ of a GET theory. ### 5.3 Existing research about the similarity Considering the mentioned problem to obtain a Galilean invariant theory it is no wonder that the usual Lagrange formalisms for non-relativistic fluid dynamics in Euler coordinates (as far as considered in Wagner ) do not fit into our scheme. Remarkably, to use the three-dimensional Einstein-Hilbert Lagrangian to describe dislocations has been proposed by Malyshev . On the other hand, it is widely acknowledged in the condensed matter community that phonons in various matter move in an effective Lorentz metric $`g_{\mu \nu }`$ which is usually curved. Various aspects have been considered here. Katanaev and Volovich compare wedge dislocation with cosmic strings. See also Guenther . A lot of research has been related with the idea of “dumb holes” – an analog of “black holes” in acoustics. These “dumb holes” may be used to study quantum gravity effects like Unruh radiation in usual condensed matter (Unruh , Jacobson , Rosu , Visser ). The most interesting example of usual condensed matter is superfluid $`{}_{}{}^{3}He`$. Here not only a curved Lorentz metric has been identified, but also chiral fermions and non-abelian gauge fields (Volovik , Jacobson and Volovik ). Of course, the amount of research which connects condensed matter theory and fundamental field theory is much greater. As noted by Wilczek , “the continuing interchange of ideas between condensed matter and high energy theory, through the medium of quantum field theory, is a remarkable phenomenon in itself. A partial list of historically important examples includes global and local spontaneous symmetry breaking, the renormalization group, effective field theory, solitons, instantons, and fractional charge and statistics.” ## 6 Comparison with RTG There is also another theory with almost the same Lagrangian – the “relativistic theory of gravity” (RTG) proposed by Logunov et al. . In this theory, we have a Minkowski background metric $`\eta _{\mu \nu }`$. The Lagrangian of RTG is $$L=L_{Rosen}+L_{matter}(g_{\mu \nu },\psi ^m)\frac{m_g^2}{16\pi }(\frac{1}{2}\eta _{\mu \nu }g^{\mu \nu }\sqrt{g}\sqrt{g}\sqrt{\eta })$$ If we identify the Minkowski coordinates in RTG with the preferred coordinates in GET, the Lagrangians are equivalent as functions of $`g_{\mu \nu }`$ for the following choice of constants: $`\mathrm{\Lambda }=\frac{m_g^2}{2}<0`$, $`\mathrm{\Xi }=\eta ^{11}\frac{m_g^2}{2}>0`$, $`\mathrm{{\rm Y}}=\eta ^{00}\frac{m_g^2}{2}>0`$. In this case, the equations for $`g^{\mu \nu }`$ coincide. The harmonic equation for the Minkowski coordinates hold in RTG . As a consequence, the equations of the theories coincide. Nonetheless, the Euler-Lagrange equations are not all. In above theories we have additional restrictions related with the notion of causality – causality conditions. In GET, causality is related with the Newtonian background – the preferred time $`T(x)`$ should be a time-like function. This is equivalent to the condition $`\rho >0`$. In RTG, causality is defined by the Minkowski background. The light cone of the physical metric $`g_{\mu \nu }`$ should be inside the light cone of the background metric $`\eta _{\mu \nu }`$. Once RTG is a special-relativistic theory, it is also incompatible with the EPR criterion of reality and Bohmian mechanics. This question should be considered as the most serious difference. There is also a difference in the quantization concept. RTG suggests to apply standard quantum field theory on a Minkowski background, while GET suggests to understand quantum field theory as an effective field theory. The GET prediction about the cutoff length depends on the interpretation of $`g^{00}\sqrt{g}`$ as the density of the ether and is not Lorentz-invariant. RTG has a completely different metaphysical background. Therefore, RTG has a completely different justification of the Lagrangian. While such metaphysical differences are often considered to be unimportant in physics, we do not agree. Metaphysical interpretations and esthetic feelings often influence preferences for theories. Because the simplicity and beauty of the explanation of the Einstein equivalence principle is one of the main advantages of GET, this question should not be underestimated. ## 7 Comparison with GR with four dark matter fields There is also another theory with the same Lagrangian – GR with four scalar “dark matter fields” $`X^\mu (x)`$<sup>4</sup><sup>4</sup>4Kuchar has considered similar scalar fields in GR as “clock fields”.. Let’s denote it as GRDM. The Lagrangian is $$L_{GRDM}=(8\pi G)^1\gamma _{\mu \nu }X_{,\alpha }^\mu X_{,\beta }^\nu g^{\alpha \beta }\sqrt{g}+L_{GR}(g_{\mu \nu })+L_{matter}(g_{\mu \nu },\phi ^m)$$ and therefore formally equivalent to GET. But we have completely forgotten the physical meaning of the fields $`X^\mu (x)`$ as coordinates. In GRDM, they are really only scalar fields. We have to consider here the usual GR energy conditions. They require the following signs: $`\mathrm{{\rm Y}}<0`$ and $`\mathrm{\Xi }>0`$. There is another difference between GET and GRDM which is essential and important to understand. In GET, we have additional global restrictions: * First, the fields $`X^\mu (x)`$ are global coordinates in GET. In GRDM, there will be many solutions where the dark matter fields do not define a global system of coordinates. Moreover, it will be even the typical solution of GRDM. Indeed, solutions which define global coordinates have unusual boundary conditions. Moreover, complete classes of solutions are excluded: all solutions with non-trivial topology are forbidden. * Second, the coordinate $`T(x)=X^0(x)`$ is a global time-like function. This is equivalent to $`\rho >0`$. Again, a whole class of solutions of GRDM is excluded: all solutions with closed time-like curves. * We have also other, unusual boundary conditions for the fields $`X^\mu (x)`$: their boundary values go to infinity. These properties do not follow from the Euler-Lagrange equations. Instead, we have to remember that the original axioms are axioms about an ether in a Newtonian space-time. The Lagrange formalism with the “fields” $`X^\mu (x)`$ is only derived, not fundamental. Therefore, it is in no way a weakness of GET that the additional global restrictions do not follow from the GET equations. Instead, this proves additional predictive power of GET in comparison with GRDM. Indeed, if we observe a solution of GRDM which does not fulfill the additional global restrictions we have falsified GET, but not GRDM. Thus, in GET we have additional possibilities for falsification, therefore, higher empirical content. This difference is of principal, conceptual character – similar global restrictions are impossible in general-relativistic theories. Therefore, it helps to understand the difference between general-relativistic theories and theories with preferred frame. We consider this question also in appendix B. ## 8 Comparison with General Relativity Let’s consider now the differences between the predictions of GET and GR itself. There are, first, the differences between GET and the variant of GR with four dark matter fields we have already considered. This restricts GET to global hyperbolic solutions with trivial topology, moreover, of a special type – with global harmonic coordinates and global harmonic time-like function. ### 8.1 Dark matter and energy conditions The other part of the difference between GET and GR can be understood as the difference between GRDM and GR. First, the additional terms define “dark matter” in the sense that the scalar fields $`X^\mu (x)`$ do not interact with usual matter. To understand the most interesting property of this new type of “dark matter” we have to consider the energy conditions. GET does not yet relate energy conditions with general properties of the ether. But because GR does not provide an explanation too, this is unproblematic. We can introduce them – as in GR – as additional, yet unexplained, properties of the matter Lagrangian. But if we introduce it in this way, as a property of the matter fields, the energy conditions do not restrict the sign of $`\mathrm{\Xi }`$ and $`\mathrm{{\rm Y}}`$. Therefore, without violating the GET version of the energy condition, it is possible to set $`\mathrm{{\rm Y}}>0`$. And, because of the interesting predictions which follow from this choice, we really set $`\mathrm{{\rm Y}}>0`$. From point of view of GRDM, this choice violates all energy conditions. Therefore, GRDM contains a very special, strange type of “dark matter” which violates all energy conditions of GR. Therefore, all general theorems about GR which use various energy conditions fail. Especially the theorems about big bang and black hole singularities fail. ### 8.2 Homogeneous universe: no big bang singularity Let’s consider at first the homogeneous universe solutions of the theory. Because of the Newtonian background frame, only a flat universe may be homogeneous. Thus, we make the ansatz: $$ds^2=d\tau ^2a^2(\tau )(dx^2+dy^2+dz^2)$$ Now, we see that in this ansatz the spatial coordinates $`x^i`$ are already harmonic. It remains to find the harmonic time. The equation for harmonic time is $`dT/d\tau =1/a^3`$. The metric in harmonic coordinates is therefore: $$ds^2=a^6(t)dT^2a^2(t)\delta _{ij}dX^idX^j$$ Note that $`\rho =g^{00}\sqrt{g}=1`$, thus, in this ansatz the ether has constant density, and the universe does not expand. The observable expansion is an effect of shrinking rulers. In GR this would be only one of the possible interpretations, without physical importance. Instead, in GET this is the preferred interpretation because of symmetry reasons. Thus, if the universe is homogeneous, the universe does not expand, but our rulers are shrinking. But, as in GR, the global universe looks like expanding. Below we use standard relativistic language and usual proper time $`\tau `$ (that means,$`\dot{a}=a/\tau `$). Using some matter with $`p=k\epsilon `$ we obtain the equations ($`8\pi G=c=1`$): $`3(\dot{a}/a)^2`$ $`=`$ $`\mathrm{{\rm Y}}/a^6+3\mathrm{\Xi }/a^2+\mathrm{\Lambda }+\epsilon `$ $`2(\ddot{a}/a)+(\dot{a}/a)^2`$ $`=`$ $`+\mathrm{{\rm Y}}/a^6+\mathrm{\Xi }/a^2+\mathrm{\Lambda }k\epsilon `$ The $`\mathrm{{\rm Y}}`$-term influences only the early universe, its influence on later universe may be ignored. But, if we assume $`\mathrm{{\rm Y}}>0`$, the qualitative behavior of the early universe changes in a remarkable way. We obtain a lower bound $`a_0`$ for $`a(\tau )`$ defined by $$\mathrm{{\rm Y}}/a_0^6=3\mathrm{\Xi }/a_0^2+\mathrm{\Lambda }+\epsilon $$ The solution becomes symmetrical in time. Therefore, before the big bang there was a big crush, and the whole story can be named big bounce. For some simple situations, analytical solutions are possible. For example, if $`\epsilon =\mathrm{\Xi }=0,\mathrm{{\rm Y}}>0,\mathrm{\Lambda }>0`$ we have the solution $$a(\tau )=a_0\mathrm{cosh}^{1/3}(\sqrt{3\mathrm{\Lambda }}\tau )$$ ### 8.3 Is there independent evidence for inflation theory? Now, such a big bounce scenario solves the problems of the big bang scenario with the small horizon. For the description of these problems and their current solution in inflation theory we follow Primack . There are two such problems with a small horizon: First, “the angular size today of the causally connected regions at recombination ($`p^++e^{}H`$) is only $`\mathrm{\Delta }\theta 3^o`$. Yet the fluctuation in the temperature of the cosmic background radiation from different regions is very small: $`\mathrm{\Delta }T/T10^5`$. How could regions far out of causal contact have come to temperatures that are so precisely equal? This is the ‘horizon problem’.” (p.56) Even more serious seems the following problem: In the standard hot big bang picture, “the matter that comprises a typical galaxy, for example, first came into causal contact about a year after the big bang. It is hard to see how galaxy-size fluctuations could have formed after that, but even harder to see how they could have formed earlier” (p.8). Last not least, there is the “flatness problem”. In GR, the assumption that the universe is flat does not seem to be natural. But for a curved universe, the initial curvature has to be extremely small in comparison to a natural dimensionless constant for curvature. Now, these three problems seem sufficient to rule out the standard Big Bang model without inflation. But all three problems are solved in GET without inflation: the two variants of the horizon problem are solved because the horizon in a universe with big bounce is much larger, if not infinite. And the flat universe is certainly preferred as the only homogeneous universe, therefore, there is no flatness problem. Therefore, it seems reasonable to question the necessity for inflation in GET cosmology. There are some other problems solved by inflation: that it “dilutes any preceding density of monopoles or other unwanted relics”, and predictions about a “nearly constant curvature spectrum $`\delta _H=`$ constant of adiabatic fluctuations” (p.59). If these will be serious problems for a GET universe without inflation is hard to say. If such “relics” are really necessary because of particle theoretical reasons, it seems possible to use a large enough $`\mathrm{{\rm Y}}`$. Then the critical temperature which causes the creation of the various “relics” may not be reached during the Big Bounce. What GET allows to predict about the spectrum of adiabatic fluctuation will be a question for future research, but it seems not unreasonable to assume that the simplest imaginable spectrum – the constant one – may be compatible with GET. Another question is if inflation is a necessary consequence of particle theory. This seems to be not the case. To obtain inflation, we have to make non-trivial assumptions about this phase transition.<sup>5</sup><sup>5</sup>5“In the first inflationary models, the dynamics of the very early universe was typically controlled by the self-energy of the Higgs field associated with the breaking of a Grand Unified Theory (GUT) into the standard 3-2-1 model: $`GUTSU(3)_{color}\left[SU(2)U(1)\right]_{electroweak}`$. … Guth (1981) initially considered a scheme in which inflation occurs while the universe is trapped in an unstable state (with the GUT unbroken) on the wrong side of a maximum in the Higgs potential. This turns out not to work … The solution in the ‘new inflation’ scheme … is for inflation to occur after barrier penetration (if any). It is necessary that the potential of the scalar field controlling inflation (‘inflaton’) be nearly flat (i.e. decrease very slowly with increasing inflaton field) for the inflationary period to last long enough. This nearly flat part of the potential must then be followed by a very steep minimum, in order that the energy contained in the Higgs potential be rapidly shared with the other degrees of freedom (‘reheating’). A more general approach, ‘chaotic’ inflation, has been worked out … However, … it is necessary that the inflaton self-coupling be very small … This requirement prevents the Higgs field from being the inflaton.” , p. 57. This consideration suggests that inflation is in no way a necessary consequence of the phase transition related with a GUT. Particle theory does not give independent evidence in favor of inflation. Thus, it seems that GET cosmology with $`\mathrm{{\rm Y}}>0`$ is a viable theory without inflation, while GR requires inflation. The existing evidence for a hot state of the universe may be used to obtain upper bounds for $`\mathrm{{\rm Y}}`$. ### 8.4 A new dark matter term The influence of the $`\mathrm{\Xi }`$-term on the age of the universe is easy to understand. For $`\mathrm{\Xi }>0`$ it behaves like homogeneously distributed dark matter with $`p=(1/3)\epsilon `$. It influences the age of the universe. A similar influence on the age of the universe has a non-zero curvature in GR cosmology. It seems not unreasonable that a non-zero value for $`\mathrm{\Xi }`$ may be part of the solution of the dark matter problem. According to Primack , there seems to be a large amount of “cold dark matter” (CDM), but this is not sufficient to fit the data. The models favored in this paper have some additional homogeneous component: some “hot dark matter” part (CHDM) or a non-zero cosmological constant ($`\mathrm{\Lambda }`$CDM). Now, the “dark matter” term proposed here is also homogeneous, and something between homogeneous “hot dark matter” and a cosmological constant. Thus, the $`\mathrm{\Xi }`$-term defines a reasonable candidate for dark matter. Current observation seems to favor $`\mathrm{\Xi }>0`$. ### 8.5 Stable frozen stars instead of black holes Let’s consider now spherical symmetric stable solutions. Of course, for symmetry reasons, we want to have static preferred coordinates too. For the preferred coordinates $`X^i`$ the metric should be harmonic. Of course, we describe this metric using the “preferred radius” $`r=\sqrt{\delta _{ij}X^iX^j}`$. Fortunately, there is a simple general formula for the harmonic metric. For a given function $`m(r),0<m<r`$, the metric $$ds^2=(1\frac{mm/r}{r})(\frac{rm}{r+m}dt^2\frac{r+m}{rm}dr^2)(r+m)^2d\mathrm{\Omega }^2$$ is harmonic in $`X^i`$. For constant m, this formula reduces to the Schwarzschild metric in harmonic coordinates. Therefore, m(r) defines the (harmonic) Schwarzschild radius of the mass inside the radius r. Now this general solution of the harmonic equation may be used to construct various partial solutions for special matter equations. We start with an arbitrary distribution of mass m(r) with $`0<m<r,m/r0`$. The Einstein equations define $`\epsilon (r)`$ and $`p(r)`$, and we obtain a solution for some material law which depends on the radius: $`p=k(r)\epsilon `$. As a simple example, let’s consider the ansatz $`m(r)=(1\mathrm{\Delta })r`$. We obtain $`ds^2`$ $`=`$ $`\mathrm{\Delta }^2dt^2(2\mathrm{\Delta })^2(dr^2+r^2d\mathrm{\Omega }^2)`$ $`0`$ $`=`$ $`\mathrm{{\rm Y}}\mathrm{\Delta }^2+3\mathrm{\Xi }(2\mathrm{\Delta })^2+\mathrm{\Lambda }+\epsilon `$ $`0`$ $`=`$ $`+\mathrm{{\rm Y}}\mathrm{\Delta }^2+\mathrm{\Xi }(2\mathrm{\Delta })^2+\mathrm{\Lambda }p`$ Now, in GR we obtain only the trivial solution $`\epsilon =p=0`$. Once the cosmological constants are sufficiently small, nothing changes for moderate values of $`\mathrm{\Delta }`$. But for sufficiently small $`\mathrm{\Delta }1`$ the situation changes – we obtain a stable solution $`p=\epsilon =\mathrm{{\rm Y}}\mathrm{\Delta }^2`$. This is a stable star with a radius very close to the Schwarzschild radius, with time dilation $`\mathrm{\Delta }^1=\sqrt{\epsilon /\mathrm{{\rm Y}}}M^1`$ for a frozen star of mass M. It is seems obvious that this is not a special property of this solution, but a rather general effect. Even a very small $`\mathrm{{\rm Y}}`$-term becomes important close enough to the horizon size and allows to obtain stable solutions. For a collapsing star this term defines a counter-force which stops the collapse immediately before horizon formation and leads to a subsequent explosion. This explosion does not follow immediately, because near the bounce the movement is time-dilated too. Thus, we obtain very interesting differences for the gravitational collapse. For the outside observer, we can fit the GR predictions making $`\mathrm{{\rm Y}}`$ small enough. But even for arbitrary small $`\mathrm{{\rm Y}}>0`$ we have remarkable qualitative differences – there is no region “behind the horizon”, no singularity, and every infalling observer can observe this difference. ## 9 General-relativistic quantization problems The quantization of gravity is usually considered as one of the major problems of fundamental physics. But, it seems, this problem should be named instead “quantization of general relativity”. Indeed, Butterfield and Isham note that “… most workers would agree on the following … diagnosis of what is at the root of most of the conceptual problems of quantum gravity. Namely: general relativity is not just a theory of the gravitational field – in an appropriate sense, it is also a theory of spacetime itself; and hence a theory of quantum gravity must have something to say about the quantum nature of space and time.” Now, in GET the theory of gravity is not a theory of space-time itself, instead, it is a theory of a medium in a classical Newtonian space-time. This problem is, obviously, much easier – the most serious problems simply disappear. On the other hand, it is much less interesting – we do not learn anything new about “the quantum nature of space and time”, instead, the classical Newtonian space-time defines the fixed stage for quantization. While the Newtonian background is very simple, it is also not very interesting and remains as unexplained as in Newton’s theory. The really hard, conceptual, interesting problems of relativistic quantum gravity disappear into nothing. What remains seem to be only a few technical problems as complex as the quantization of usual condensed matter. Once these conceptual problems disappear, they can be considered as additional fundamental support for GET and are therefore worth to be considered in this context. In appendix D we consider in more detail a problem related with superposition of gravitational fields which I have named the “scalar product problem”. This problem has several nice properties: it suggests a simple solution – a fixed space-time background which is common for different gravitational fields. Moreover, it is based on an interesting quantum observable – a transition probability. And this transition probability may be computed in the non-relativistic limit – multi-particle Schrödinger theory. The notorious “problem of time” is mainly a conceptual problem. It appears if we make a deliberate theoretical decision: that the time measured with clocks – the time of general relativity – has to be unified with the notion of time of quantum mechanics. We discuss these metaphysical questions in § C. Nonetheless, some other well-known problems of quantum GR which do not exist in quantum GET are also worth to be considered: the problem of causality, and the information loss problem. ### 9.1 Causality The problem of time is also closely related with causality: “General relativity accustoms us to the ideas that (i) the causal structure of spacetime depends on the metric … and (ii) the metric and causal structure are influenced by matter … In general relativity, these ideas are ‘kept under control’ in the sense that in each model, there is of course a single metric tensor $`g_{\mu \nu }`$, representing a single metric and causal structure. But once we embark on constructing a quantum theory of gravity, we expect some sort of quantum fluctuations in the metric, and so also in the causal structure. But in this case, how are we to formulate a quantum theory with a fluctuating causal structure?” This conceptual problem has also technical aspects. “For example, a quantum scalar field satisfies the micro-causal commutation relations $$[\widehat{\varphi }(X),\widehat{\varphi }(Y)]=0$$ whereby fields evaluated at space-like separated spacetime points commute. However, the concept of two points being space-like separated has no meaning if the spacetime metric is probabilistic or phenomenological. In the former case, the most likely scenario is that \[the commutator\] never vanishes, thereby removing one of the foundations of conventional quantum field theory.” ### 9.2 Information loss problem Another problem which disappears is the “information loss problem” proposed by Hawking . The problem is that the black hole contains information. But the Hawking radiation cannot take away this information because it is determined only by the geometry of the black hole outside the horizon, and the black hole has no hair that records any detailed information about the collapsing body. The key constraint comes from causality – once the collapsing body is behind the horizon, it is incapable of influencing the radiation. Now, suppose the black hole evaporates. That means, the black hole has been replaced by the radiation completely. It is a familiar fact of life that information is often lost in practice. But here the information is lost in principle. It seems that an initially pure state becomes after evaporation a mixed state. And this is in contradiction with the fundamental principles of quantum mechanics. Preskill in a review of the problem writes that initially he “was inclined to dismiss Hawking’s proposal as an unwarranted extrapolation from an untrustworthy approximation”. But as “I have pondered this puzzle, it has come to seem less and less likely to me that the accepted principles of quantum mechanics and relativity can be reconciled with the phenomenon of black hole evaporation.” Now, it is hard for me to judge about the seriousness of this problem. Personally I’m convinced that the correct black hole evaporation scenario in GR is different from the usually accepted one. In my opinion it is the scenario of black hole evaporation proposed by Gerlach . In this scenario, no black hole horizon is formed. As far as I understand, relativists do not like this scenario because it seems to prefer the coordinates of the outside observer. But I don’t think this argument is justified – the preference is predefined by the preference for the Minkowski vacuum state in the initial situation before the collapse. In Gerlach’s scenario we have no information loss problem because no horizon is formed. Anyway, in GET the information loss problem disappears together with the black holes. We have stable frozen stars which do not radiate Hawking radiation once they have reached a stable state. We also have not to be afraid of some similar situations – there is always a global absolute time, and this global time has to be used in quantum GET to define an unitary evolution. ## 10 Atomic ether theory With the Newtonian background only the conceptual problems related with relativistic space-time quantization disappear. The problem with non-renormalizability remains. But there is a natural solution for this problem in the context of a condensed matter theory – an atomic hypothesis. Therefore, we assume that our medium has an atomic structure. This leads to an explicit cutoff. The regularization becomes physical. The concept of gravity as an effective field theory is well-known and goes back to Sakharov . <sup>6</sup><sup>6</sup>6Jegerlehner notes that ideas that “the relationship between bare and renormalizes parameters obtains a physical meaning … are quite old and in some aspects are now commonly accepted among particle physicists” . Weinberg describes this as “the present educated view on the standard model, and of general relativity, … that these are leading terms in effective field theories” . One property of this widely accepted effective field theory picture is that it makes a certain assumption about the cutoff length. It is assumed to be the Planck length $`a_P10^{33}`$cm. This property has even used to name this concept: “Planck ether”, “Planck solid” or “Planck condensed matter” . But this is de facto the only property of the “Planck ether” which is known. <sup>7</sup><sup>7</sup>7“The curvature of space-time is relevant and special relativity is modified by gravitational effects. One expects a world which exhibits an intrinsic cutoff corresponding to the fundamental length $`a_P10^{33}`$cm. But not only Poincare invariance may break down, also the laws of quantum mechanics need not hold any longer at $`\mathrm{\Lambda }_P`$.” The “atomic ether hypothesis” differs from this concept in everything except the fact that current field theory should be replaced by another one below some cutoff. First, a well-defined space-time concept has been fixed – the classical Newtonian background with Euclidean space, absolute time and classical causality. Moreover we have well-defined conservation laws. We have also fixed the quantization concept – classical canonical quantization for a theory with a discrete number of steps of freedom. Momentum quantization is part of this concept. Moreover, the number of ether atoms will not vary in this scheme too. Thus, we fix an extremely simple, very special class of underlying microscopic theories. <sup>8</sup><sup>8</sup>8It may be argued that we are unable to do experiments in this domain, therefore, it is also unreasonable to make a hypothesis about nature in this domain. All we can do is to describe their universality class. Nonetheless, even if we are unable to do experiments, we can use general principles as Ockham’s razor to make a choice. Moreover, there is no need to make a definite choice. A situation where we have several metaphysically completely different models which are all in ideal agreement with experiment would be satisfactory too – it gives a clear account about the boundaries of scientific research. Another property of atomic ether theory is especially interesting: the interpretation of the “ether density” $`\rho =g^{00}\sqrt{g}`$ as the number of “ether atoms” per volume. This leads to an interesting prediction for the cutoff: $$\rho (x)V_{cutoff}=1.$$ The point is that this prediction is different from the usual Planck length $`a_P`$. This can be illustrated with the example of the “homogeneous universe” solution (see § 8.2). In this solution, $`\rho `$ remains constant in time. Therefore, the cutoff length remains constant in time too. On the other hand, our rulers shrink. Thus, the cutoff length is not constant in our cm scale, it cannot be the Planck length. From point of view of our rulers, the cutoff length is expanding. For $`\mathrm{{\rm Y}}<0`$ or small enough $`\mathrm{{\rm Y}}`$ the cutoff was below Planck scale in the past, for $`\mathrm{\Lambda }>0`$ or small enough $`\mathrm{\Lambda }`$ it will be greater than Planck scale, but even greater than the cm scale, in future. Thus, in this case we will be able to observe in a far away cosmological future the effects of the atomic ether. ## 11 Canonical atomic ether quantization Ether theory suggests to solve the ultraviolet problems with explicit, physical regularization, based on the idea of an atomic ether. This idea leads to the following canonical atomic ether quantization scheme: 1. First, we need the full continuous ether theory, that means, a continuous “theory of everything” (TOE) which describes the complete ether. This in no way implies that we need some “grand unification” to obtain this TOE. There is no requirement that there should be only one unified force. We can as well try to start with GET + SM. 2. Then, we have to find an atomic ether model which gives the TOE in the large scale limit. There are obviously many of them, because the large scale limit fixes only the universality class. This uncertainty remains as long as there is no experimental evidence in the short distance domain. Therefore simple standard models would be sufficient. The possibly critical problem is to prove that the large scale limit of the discrete model is indeed the original theory. 3. Then, this classical atomic ether theory has to be quantized in the canonical way. If the atomic model is close enough to usual atomic models of usual condensed matter, this step will be unproblematic. Note that in this case we can apply also Bohmian mechanics. Once this step has been finished, we have quantized gravity. 4. Now, for actual quantum computations it is necessary to derive the large scale limit of the quantum atomic ether theory, which will be quantum field theory. Note that this step is already purely phenomenological and not necessary for the theory itself. That mean, problems which appear in this last step are problems of large scale approximations in classical quantum theory, but not fundamental problems of quantum gravity. Now, for this quantization program we have to distinguish different problems. The first problem is if this scheme works at all. Here, we are in a very good situation. We have an example in reality where the the whole quantization scheme is realized – superfluid $`{}_{}{}^{3}HeA`$. In the theory of $`{}_{}{}^{3}HeA`$, we obtain in the large scale limit the most important ingredients of the current standard model. Volovik writes: “In this sense the superfluid phases of $`{}_{}{}^{3}He`$, especially $`{}_{}{}^{3}HeA`$, are of most importance: the low-energy degrees of freedom in $`{}_{}{}^{3}HeA`$ do really consist of chiral fermions, gauge fields and gravity” . Thus, if we do not have an (uncommon among physicists) desire for mathematical rigor, for at least one field theory with gravity, fermions and gauge fields an atomic ether quantization scheme works. Moreover, it works in reality. This is very important – the existence of such a model in reality gives certainty that the program may be realized, and it provides suggestions how this has to be done. ### 11.1 Regularization using a moving grid Let’s see how this works on the example of our second step – the derivation of an atomic model for a given continuous ether theory. Without the ether interpretation, it would be natural to try to regularize the theory with a regular lattice, following lattice gauge theory. Based on the space-time interpretation, we would try to develop a discrete variant of geometry – something like the Regge calculus or dynamical triangulations . The ether interpretation suggests something different – a discretization which remains as close as possible to the real atomic grid. That means, we do not have to use a regular, static lattice. Instead, we have to use a grid with the following properties: * The grid node density is the “ether density” $`\rho `$: $$\underset{x_kV}{}1_V\rho d^3x$$ * The grid moves, with “ether velocity” $`v^i(x_k)`$; Without additional information about the other material properties of the ether we cannot say anything reasonable about their discretization. But in the case of the Minkowski space-time, this prescription reduces to a homogeneous, static grid. Therefore, we can use the lectures of lattice gauge theory to understand how to discretize them. Let’s consider shortly some details: as a first step to obtain an atomic model we have to switch from Euler (local) coordinates to Lagrange (material) coordinates. Once GET is defined by a Lagrange formalism, the first interesting problem is if this transformation is possible in the Lagrange or Hamilton formalism too. This is known to be possible in hydrodynamics. In the Hamilton formalism this can be done with a canonical transformation , , . In the following we assume that this is possible in GET too. The next important step is the discretization. Here, it is useful to have in mind that we want to quantize the theory later. Therefore, to be able to use canonical quantization, we need a Lagrange or Hamilton formalism for the discrete theory. For this purpose, it is not reasonable to discretize the equations them-self. Instead, it is much more reasonable to discretize the Lagrange function and to define the discrete equations as Euler-Lagrange equations for the discrete Lagrangian. The usual method to obtain a discrete function on a grid is the finite element method. In this method, we define functions on the grid as a subspace of the space of all functions. In the simplest case, this is the space of piecewise linear function on the simplices. These functions are uniquely defined by their function values on the grid nodes: $`f(x_k)=f_k`$. This defines an embedding of the grid functions into the space of all functions. In the other direction, we can use orthogonal projection of this subspace to define the discrete image of a continuous function. Thus, the function values in the nodes $`f_k`$ are not defined by the function values of the original continuous functions, but by integral formulas: $`f_k=f(x)\chi _k(x)𝑑x`$, where $`\chi _k(x)`$ is the piecewise linear function defined by $`\chi _k(x_l)=\delta _{kl}`$. In our case, we have an interesting modification of this method: the density $`\rho (x)`$ should not be described by a variable grid function $`\rho _i`$. Instead, an integral containing $`\rho `$ should be interpolated on the grid in another way. The simplest way would be $$_Af(x)\rho (x)𝑑x\underset{x_kA}{}f_k$$ Therefore, in the discrete Lagrange formalism the density simply disappears. Now, there are various variants of this method, and which is the best one depends on the material properties of the ether. Therefore, further specification of the scheme in this general context does not seem to be justified and has to be left to future research. ### 11.2 Constraints and conservation laws in a moving grid Now, it is reasonable to ask about the advantage of this type of discretization, for example in comparison with a regular, static lattice. As far, the only argument was that this looks more natural from point of view of the atomic ether interpretation. But, if we look at the remaining problems, we observe that they become essentially simplified. Indeed, once we have a well-defined discrete Lagrange formalism, the most serious remaining quantization problem are constraints. Unfortunately, without specification of the material properties we cannot say anything about possible constraints related with these material properties. But there are well-known constraints in GR, and as we have seen in § 3.2, they remain to be constraints in GET. Moreover, the conservation laws them-self are constraints too. Now, these constraints essentially change their character. The continuity equation simply disappears. It is no longer an equation of the discrete theory. The density $`\rho `$ is no longer part of the equations. Instead, the continuity equation becomes a tautology – the number of grid nodes remains constant. Moreover, the Euler equations become second order equations: the first order derivative of $`g^{0i}\sqrt{g}`$ becomes a second order derivative of the grid node position: $$_t(g^{0i}\sqrt{g})=_t(\rho v^i)m\ddot{x^i}$$ Moreover, we have seen in § 3.2 that the other constraints also may be explained by the requirement that there are no second order equations for $`\rho `$ and $`\rho v^i`$. Therefore, the constraint problem essentially simplifies. Closely related with the constraints is the question how the conservation laws are realized. The conservation of ether particles in this approach is realized automatically, as the conservation of the number of grid nodes. Therefore, we have no “quantum fluctuations” of the ether particle number. Note that this property holds in the fundamental, atomic ether theory. We do not make any claim about the large scale quantum field theory approximation and the behavior of a field operator $`\widehat{\rho }(x)`$ which fulfills an operator version of the continuity equation. This approximation is nothing we have to care as long as we consider fundamental questions. ### 11.3 Universality Now, considering the previously discussed program, it seems to suggest that almost every ether theory may be quantized in this way. But a completely different question is how a typical ether theory looks like. For this question, the consideration of long distance universality is important. The first point of long distance universality is that very different atomic theories can have the same large scale limit. “Long distance universality is a well-known phenomenon from condensed matter physics, where we know that a ferromagnet, a liquid-gas system and a super-conductor may exhibit identical long range properties (phase diagram, critical exponents, etc.)” . This point is important for the justification of the “moving grid” method – it suggests that it is not very meaningful to search for the “true” atomic theory without experimental evidence in this domain. All we can do is to search for a theory which is sufficiently simple and natural in comparison with their competitors. Even without experimental evidence Ockham’s razor may be used to choose between theories. The other side of large scale universality is that the theories which appear as large scale approximations have some very typical properties. Existing research in this domain has already given important and interesting results: “The extraction of the leading low energy asymptote is equivalent to the requirement of renormalizability of S-matrix elements, and this has been shown to be necessarily be a non-Abelian gauge theory which must have undergone a Higgs mechanism if the gauge bosons are not strictly massless. … only a renormalizable field theory can survive as a tail, the possible renormalizable theories on the other hand are known and are easy to classify” . At a first look, this seems to be in conflict with our atomic ether quantization scheme. There seems to be no point where similar restrictions appear. But there is no contradiction – these are simply different questions. One question is if a quantization in this way is possible in principle. Another question is if a theory has the typical properties of a large scale limit of an atomic ether theory. The answer for the first question may be very well a positive one, but the related atomic ether theory may require a very strange conspiracy of their coefficients. But, if we have seen, for the basic ingredients of the SM – fermions and gauge fields – the situation is very nice. It is well understood in theory why such renormalizable theories appear as large scale limits of a typical atomic ether theory, and we have observed them in reality in condensed matter, in $`{}_{}{}^{3}HeA`$. Considering all these facts, it seems likely that this scheme works, and that problems which appear on this way may be solved. Instead, the problem of canonical GR quantization we discuss in § D suggests that a quantization without a fixed background fails in principle. ## 12 Comparison with canonical quantization of general relativity It is interesting to compare our canonical program for GET quantization with the real way of development of quantization programs for general relativity, especially the canonical program. The point is that the progress of the canonical quantization program is much more in agreement with ether philosophy than with general relativistic philosophy. Let’s consider the different steps in the standard canonical quantization approach: ### 12.1 ADM formalism The ADM decomposition is essentially the decomposition of $`g_{\mu \nu }`$ into $`\rho ,v^i,p^{ij}`$. Therefore, it is an essential part of the GET approach. This decomposition is in obvious disagreement with relativistic philosophy. Space and time are no longer considered as a unit, they are separated. We have a special time coordinate t(x). Moreover, the general spacetime manifold becomes subdivided into a product $`S`$, that means, changes of topology in time are excluded. This modification of relativistic metaphysics is so important that the ADM formalism is often considered as a different interpretation of general relativity – geometrodynamics. In this interpretation, GR no longer describes a spacetime, but the evolution of three-dimensional geometries. Of course, the Hamilton formalism of general relativity and that of ether theory remain quite different. In GET, the Hamiltonian is not a constraint. We have equations for the preferred coordinate $`T(x)`$ used to define the foliation. And we have additional physical steps of freedom: density and velocity of the ether. ### 12.2 Tetrad and triad formalism The next important step is the introduction of the tetrad formalism. In this formalism, the metric $`g^{\mu \nu }`$ becomes derived. We have a tetrad field – four vector fields $`e_a^\mu (x)`$ which form an orthonormal basis in each point, so that $$g^{\mu \nu }=e_a^\mu e_b^\nu \eta ^{ab}.$$ This is necessary for the incorporation of fermions into general relativity. Obviously this step is also a gross violation of relativistic ideology. Indeed, according to this ideology the field $`g^{\mu \nu }`$ has already a fundamental interpretation, it defines spacetime. The other argument in favor of this procedure is that GR becomes a gauge theory, with the gauge group SO(3,1). The next step on this way is the combination of above approaches: The time coordinate of the ADM composition is used to fix the time-like tetrad vector. After this “time gauge” we have a triad – three vector fields in space, with compact gauge group SO(3). The technical problems with non-compact gauge groups are a main argument for this choice. Now, a triad field looks very natural from point of view of ether theory. If we imagine the ether as a crystal, the three triad vector fields may be considered as defining locally the orientation of the crystal structure. Thus, to introduce triad variables may be an interesting possibility for canonical GET quantization. Of course, as in the case of the ADM variables, we cannot take the formulas as they are, because we have a different Hamiltonian and different steps of freedom. Instead, to consider this triad formalism as something natural from point of view of relativistic ideology seems impossible. ### 12.3 Ashtekar variables The next step is a canonical transformation to Ashtekar variables which simplifies the constraints. There are two variants of the Ashtekar formalism: the first was a complex formalism. In this complex formalism, an additional simplification of the Hamiltonian constraint happens. For this advantage it is necessary to pay with the problem of “reality conditions”. But the problem with these reality conditions was too hard, that’s why following Barbero the real version of the Ashtekar formalism is preferred now. This seems to be a good idea from point of view of the ether approach – the real variant of the formalism clearly better fits into ether ideology in comparison with a complex formalism. The physical meaning of the Ashtekar variables is not obvious at all in the relativistic approach. On the other hand, we already know that we have to do something similar, with similar results, but with clear physical interpretation in canonical ether quantization: the transformation from Euler to Lagrange coordinates. As we have already mentioned, this is a canonical transformation, and it results in a simplification of the constraints. ### 12.4 Discrete models of geometry An important part of existing attempts to quantize gravity are discrete models. In some sense, discrete models are also not in ideal fit with classical spacetime ideology. It would be much more natural to have a continuous spacetime for all distances. But the problems with non-renormalizability suggest to use discrete regularizations. On the other hand, the consideration of discrete models is a natural part of canonical ether quantization, with clear physical motivation: an atomic ether theory. Of course, the grids used in ether theory are three-dimensional grids in a standard Newtonian space, moving in continuous time. The position of the grid nodes are steps of freedom of the ether. These steps of freedom do not exist in the purely geometrical approaches. Nonetheless, we can learn from these approaches how to discretize the geometric steps of freedom. For example, it seems quite natural to use the basic ideas of Regge calculus to discretize the pressure $`p^{ij}`$. We obtain a discretization where the pressure $`p^{ij}`$ is described by a scalar on each edge between neighbor nodes. This discretization has a natural interpretation as the force between neighbor atoms. ### 12.5 Summary The interesting observation is that, while searching for a way to quantize general relativity, most success has been reached in a direction which is in no way close to standard GR ideology: * introduction of a preferred time and a Hamilton formalism; * introduction of other variables so that the metric $`g_{\mu \nu }`$ no longer fundamental; * canonical transformations to simplify the constraints; * discretization of the theory; Instead, all these steps are quite natural in the ether approach. We have formulated most of them in our canonical ether quantization program. The steps of freedom in GET are different from the steps of freedom in canonical GR, and the formulas for the GET approach have yet to be worked out. Nonetheless, our observation suggests that the canonical ether quantization concept is on the right way. ## 13 Quantum field theory Now, the question how to quantize an ether theory is conceptually completely different from the quantization of GR. The main question we want to consider here is if this leads to differences in semi-classical QFT. In principle, the way we have to quantize continuous ether theory is to quantize a discrete atomic ether model in a canonical way and then to consider the large scale limit. Thus, we have to quantize gravity similar to the quantization of hydrodynamics by extrapolation of the microscopic theory, as done by Landau . But it has been found (Davydov ) that the same result may be obtained by canonical quantization, without using microscopic theory. Therefore, without having reasonable microscopic models, it is reasonable to apply canonical quantization to the continuous GET equations. To consider microscopic models seems necessary only for a better understanding of the way we have to regularize the infinities. For example, we can learn why the renormalization of the vacuum energy is justified. This can be seen using superfluid $`{}_{}{}^{3}He`$ as a model (Volovik ). Once we use canonical quantization, it is no wonder that we obtain the same formulas as usual in quantum field theory. Nonetheless, some remarks seem to be interesting. First, even if in our covariant formulation the preferred coordinates formally appear as fields $`X^\mu (x)`$, this does not mean that they should be quantized as scalar fields. This would be a serious misunderstanding about the purpose of the covariant formulation. Instead, the $`X^\mu `$ remain classical preferred coordinates. This is an immediate consequence of the basic idea for quantization: to quantize a microscopic atomic model in a canonical way, using classical Schrödinger theory. ### 13.1 Semi-classical quantization of a scalar field Let’s consider as an example the canonical quantization of a scalar field on a classical GET background. We have the Lagrangian $$=\frac{1}{2}\sqrt{g}(g^{\mu \nu }\varphi _{,\mu }\varphi _{,\nu }m^2\varphi ^2)$$ We have a well-defined preferred frame defined by the coordinates $`X^\mu `$, and we quantize the field in this frame. Note that canonical quantization is a very artificial procedure from point of view of general relativity - it destroys its covariance ideology. Instead, it is a very natural procedure from point of view of ether theory. Using the standard formalism of canonical quantization, we obtain $$\pi =\frac{}{\varphi _{,0}}=\widehat{g}^{0\mu }\varphi _{,\mu }$$ $$=\pi \varphi _{,0}=\frac{1}{2}(\widehat{g}^{00})^1(\pi \widehat{g}^{0i}\varphi _{,i})^2\frac{1}{2}\widehat{g}^{ij}\varphi _{,i}\varphi _{,j}+\frac{m^2}{2}\varphi ^2\sqrt{g}$$ Note that these expressions look beautiful in the original ether variables too: $$\pi =\rho \varphi _{,0}+\rho v^i\varphi _{,i}$$ $$=\frac{1}{2}(\rho ^1\pi ^22\pi v^i\varphi _{,i}p^{ij}\varphi _{,i}\varphi _{,j})+\frac{m^2}{2}\varphi ^2\sqrt{\rho |p^{ij}|}$$ As we see, our ADM-like decomposition is in good agreement with the canonical formalism. We define now $`\varphi `$ and $`\pi `$ as operators with the standard commutation rules ($`\mathrm{}=1`$): $$[\varphi (x),\pi (y)]=i\delta (xy)$$ As we see, this definition does not depend on the gravitational field. This is an important observation. The background space, its affine symmetry, the related Hilbert space for the field $`\varphi (x)`$, the commutation relations and the algebra of observables on this Hilbert space do not depend on the gravitational field. This does not seem to be important in semi-classical theory, but it becomes very important if we consider superpositions of gravitational fields (see appendix D). In this case, the definition of the Hilbert space may be used as it is, and scalar products between states defined for different gravitational fields are well-defined. This is a very important difference between quantization of GR and GET. In GR, the spacetime points and therefore the Hilbert spaces for the fields $`\phi (x)`$ have no independent meaning. ### 13.2 Particle operators and vacuum state One of the main lectures of quantum field theory is that the fundamental object are the fields, not the particles. The notion of particles is derived, secondary. <sup>9</sup><sup>9</sup>9“In its mature form, the idea of quantum field theory is that quantum fields are the basis ingredients of the universe, and particles are just bundles of energy and momentum of the fields” . What “quantum field theory uniquely explains is the existence of different, yet indistinguishable, copies of elementary particles. Two electrons anywhere in the Universe, whatever their origin and history, are observed to have exactly the same properties. We understand this as a consequence of the fact that both are excitations of the same underlying ur-stuff, the electron field. The electron field is thus the primary reality” . GET does not question this insight. Instead, in the canonical GET quantization scheme this becomes exceptionally obvious. The classical continuous ether is described by continuous fields – properties of the ether. The fields are fundamental. Their description does not depend on the gravitational field – as we have seen, the Hilbert space for the quantum field $`\phi (x)`$ is defined independent of the gravitational field. On the other hand, the notion of particles and the vacuum state do not appear in a gravity-independent way. For the vacuum state we have a natural definition: it is the state with minimal energy. But the Hamilton operator depends on the gravitational field, therefore, the definition of the vacuum state and the notion of particles too. In the case of a constant metric $`g^{\mu \nu }`$ particle operators are defined by the formulas: $`\varphi _k`$ $`=`$ $`{\displaystyle e^{ikx}\varphi (x)𝑑x}`$ $`\pi _k`$ $`=`$ $`{\displaystyle e^{ikx}\pi (x)𝑑x}`$ $`a_k^+`$ $`=`$ $`{\displaystyle \frac{1}{\sqrt{2\omega _k}}}(\pi _ki(\widehat{g}^{0i}k_i\omega _k)\varphi _k),`$ $`a_k`$ $`=`$ $`{\displaystyle \frac{1}{\sqrt{2\omega _k}}}(\pi _ki(\widehat{g}^{0i}k_i+\omega _k)\varphi _k),`$ $`\omega _k^2`$ $`=`$ $`\widehat{g}^{00}(\widehat{g}^{ij}k_ik_j+m^2\sqrt{g})`$ with $`H={\displaystyle 𝑑x}`$ $`=`$ $`{\displaystyle \frac{1}{2}}{\displaystyle \pi _k^2}+\omega _k^2\varphi _k^2dk`$ $`[a_k,H]`$ $`=`$ $`\omega _ka_k`$ $`a_k|0`$ $`=`$ $`0`$ Now, in the case of a non-trivial gravitational field, these particle states are no longer eigenstates of the the Hamilton operator. They interact with the gravitational field. Nonetheless, they remain to be an approximation. We can introduce wave packets: $`\varphi _{kx}`$ $`=`$ $`{\displaystyle e^{iky\sigma (yx)^2}\varphi (y)𝑑y}`$ $`\pi _{kx}`$ $`=`$ $`{\displaystyle e^{iky\sigma (yx)^2}\pi (y)𝑑y}`$ $`a_{kx}^+`$ $`=`$ $`{\displaystyle \frac{1}{\sqrt{2\omega _{kx}}}}(\pi _{kx}i(\widehat{g}^{0i}k_i\omega _{kx})\varphi _{kx})`$ $`a_{kx}`$ $`=`$ $`{\displaystyle \frac{1}{\sqrt{2\omega _{kx}}}}(\pi _{kx}i(\widehat{g}^{0i}k_i+\omega _{kx})\varphi _{kx})`$ $`\omega _{kx}^2`$ $`=`$ $`\widehat{g}^{00}(\widehat{g}^{ij}k_ik_j+m^2\sqrt{g})`$ We obtain $`[a_{kx},H]`$ $``$ $`\omega _{kx}a_{kx}`$ $`a_{kx}|0`$ $``$ $`0`$ Here the vacuum state $`|0`$ remains to be defined as the state with minimal energy, it’s expression using the local particle operators becomes an approximation. This local definition of particle is useful for comparison with existing semi-classical field theory (cf. ). In this theory, we have the problem how to define the vacuum state and the Fock space. It is usually solved by definition of a set of observers. For these observers, the vacuum state is the state where they do not observe particles. Now, a similar problem does not appear in our canonical scheme. for our scalar field we have a natural choice – the vacuum state as the state with minimal energy. But this choice may be understood in a similar way as the definition of a set of preferred observers – the observers which are in rest compared with the preferred frame. Now, our local particle operators may be interpreted as the particle operators which are important for the local observers and their particle detectors. In the vacuum state they do not observe particles. Therefore, our definition of the vacuum is in agreement with the definition of the vacuum state related with the set of preferred observers which are in rest. ### 13.3 Different representations The major technical problem in quantum field theory on a curved background $`(M,g_{\mu \nu })`$ is the existence of infinitely many unitarily inequivalent representations of the canonical commutation relations. Isham describes it in the following way: “The real problems arise if one is presented with a generic metric $`g_{\mu \nu }`$, in which case it is not at all clear how to proceed. A minimum requirement is that the Hamiltonians $`H(t)`$, or the Hamiltonian densities should be well-defined. However, there is an unpleasant possibility that the representations could be $`t`$ dependent, and in such a way that those corresponding to different values of $`t`$ are unitarily inequivalent, in which case the dynamical equations are not meaningful.” Now, GET gives all what may be wanted to prove that this does not happen: we have a simple equation for the metric (the harmonic equation), conservation of some important quantities (ether mass and momentum), we can use inequalities for $`\rho `$ and p of type $`\epsilon <\rho <R`$ (which may be interpreted as boundaries for the validity of GET) if necessary. Possibly this will be sufficient to solve this technical problem. On the other hand, this problem appears also in thermodynamics for states with different temperature – something which in reality sometimes changes in time. Therefore it would not be strange if the problem nonetheless remains. I consider it to be an artifact of the limit $`l_{cutoff}0`$. ### 13.4 Gauge field quantization Let’s consider now questions related with the quantization of gauge fields, at first the simplest case of QED in flat space. There are different well-known quantization schemes which may be used to incorporate the gauge condition. In the variant of Bjorken and Drell the gauge condition (Coulomb gauge) is incorporated into the configuration space: $$[\dot{A_i}(𝐱,t),A_j(𝐱^{},t)]=i\delta _{ij}^{tr}(𝐱𝐱^{})$$ The other possibility is to consider a large configuration space $$[\dot{A_\mu }(𝐱,t),A_\nu (𝐱^{},t)]=i\delta _{\mu \nu }\delta (𝐱𝐱^{})$$ and to incorporate the gauge condition as an additional restriction for the states: $$(_\mu A_\mu (x))_+|\mathrm{\Phi }=0$$ This scheme leads to a problem with the interpretation of the particle operators. We have $$[c_{k\mu }c_{k^{}\nu }^{}]=\eta _{\mu \nu }\delta kk^{}$$ therefore the role of $`c_{k0}`$ is reversed: $`c_{k0}`$ behaves like $`c_{kj}^{}`$. Now, there are two variants of the interpretation of these commutation relations. In the first, classical, Fermi-Dirac quantization , we accept that the Lorentz symmetry is broken in the large space: the vacuum is defined by $$c_{k0}^{}|\mathrm{\Phi }_0=c_{ki}|\mathrm{\Phi }_0=0$$ In the other, explicitly relativistic variant introduced by Gupta and Bleuler , we define the vacuum in the invariant way $$c_{k\mu }|\mathrm{\Phi }_0=0$$ and obtain an indefinite Hilbert space: $$\mathrm{\Phi }_0|c_{k0}c_{k0}^{}|\mathrm{\Phi }_0<0$$ Again, we have a conflict between relativistic symmetry and a fundamental physical principle – the definiteness of the Hilbert space. Of course, it is well-known that these differences do not lead to observable differences. Nonetheless, this particular quantization problem is further illustration of the general picture we have found in Bohmian mechanics as well as for the local energy and momentum of the gravitational field: every more fundamental description requires to break relativistic symmetry. It is obvious that in this case we prefer the definite Hilbert space. Therefore, GET suggests to reject the Gupta-Bleuler approach and to use, instead, the older, non-covariant Fermi-Dirac quantization scheme. The choice between Fermi-Dirac quantization and the scheme used by Bjorken and Drell is more complicate. We prefer the Fermi-Dirac quantization scheme because it is based on the Lorenz condition $$_\mu A^\mu =0.$$ This condition is interesting for GET because of the known analogy between gauge theory and gravity. Once GET modifies the understanding of gravity and the EEP, a similar modification of gauge symmetry would be natural. In this scenario, the Lorenz condition would be the natural candidate for a physical equation, similar to the harmonic condition in GET. It also allows a physical interpretation as a conservation law of some ether property. The incorporation of exact conservation laws into a field theory is a subtle thing: while integrals over a finite domain very in time, the integral over the whole space should be exactly conserved, without any quantum fluctuations. An atomic ether model suggests natural ways to reach this property – if conservation laws are interpreted as conservation laws for numbers of atoms, and canonical multi-particle Schrödinger theory is used to quantize the theory, this number is conserved automatically. In field theory I don’t know such a way. But it should be noted here again that if this is a problem, it is a problem of the field theory approximation and therefore not a fundamental problem of the ether approach. Field theory is only an approximation, their problems are therefore not fundamental problems, but problems of an inconsistent approximation. ### 13.5 Non-abelian gauge field quantization We do not consider here the quantization of non-abelian gauge fields. The reason is not that this seems to be very hard. It is certainly not impossible, because they appear in real condensed matter (SU(2) in superfluid $`{}_{}{}^{3}HeA`$ ). They may be justified as renormalizable theories which appear in a natural way in the large scale limit . For the development of discrete atomic ether models we can also use the large amount of experience with lattice QCD . The reason is simply that I have not considered this domain yet in sufficient detail. ### 13.6 Hawking radiation For an instationary gravitational field the vacuum state and the particle operators depend on time. Therefore, the original vacuum becomes a state with particles. Once the basic concept remains unchanged, we obtain the same results: ###### Theorem 3 Let $`g_{\mu \nu }(X,T)`$ be a background metric in preferred coordinates $`X^i,T`$ and let’s denote the set of observers in rest compared with the preferred coordinates the “preferred observers”. Then in the canonical formalism we obtain the same results for Hawking radiation as in the usual formalism for the preferred set of observers. Indeed, the formalism does not depend on the question if the background metric is a solution of GET or GR. The only difference with the standard formalism is the well-defined choice of the preferred observers in every moment of time. But this is simply the application of the general formalism to this special choice of preferred observers. But this does not mean that semi-classical GET predicts Hawking radiation similar to semi-classical GR. The equivalence holds only as long as we use the same metric $`g_{\mu \nu }`$. For usual configurations this can be done, the cosmological constants $`\mathrm{{\rm Y}}`$ and $`\mathrm{\Xi }`$ may be ignored. But for the interesting case of the gravitational collapse this is not the case. We obtain a stable “frozen star” without horizon (see §8.5). Therefore, during the collapse we obtain Hawking radiation. But once the collapse has stopped, the radiation goes away and no new radiation appears, as for stable stars in GR too. The remarkable result is that this does not depend on the actual value of $`\mathrm{{\rm Y}}>0`$. Even for very small $`\mathrm{{\rm Y}}`$ the collapsing star needs only a short time to reach the critical size of the frozen star. Once a stable state has been reached, the radiation disappears. Stable stars do not radiate. Therefore, GET predicts no Hawking radiation from frozen stars. For small enough “frozen balls” this leads to observable differences between GR and GET. In GET they remain stable and don’t evaporate. ## 14 Methodology Because of the lack of data, in the domain of quantum gravity methodological and philosophical questions become much more important than in other domains of science. In some sense, they become decisive.<sup>10</sup><sup>10</sup>10Butterfield & Isham describe this situation in the following words: “there are no phenomena that can be identified unequivocally as the result of an interplay between general relativity and quantum theory - a feature that arguably challenges the right of quantum gravity to be considered as a genuine branch of science at all! … theory construction inevitably becomes much more strongly influenced by broad theoretical considerations, than in mainstream areas of physics. More precisely, it tends to be based on various prima facie views about what the theory should look like – these being grounded partly on the philosophical prejudices of the researcher concerned … In such circumstances, the goal of a research programme tends towards the construction of abstract theoretical schemes which are compatible with some preconceived conceptual frameworks”. Does it mean that it is impossible to find agreement about methodological issues? Fortunately, the methodological concepts proposed by Rovelli are in good agreement with the method used here. The key idea of his methodology is the following: > … confidence in the insight that came with some theory, or ‘taking a theory seriously’, lead to major advances that largely extended the original theory itself. Of course, far from me suggesting that there is anything simple, or automatic, in figuring out where the true insights are and in finding the way of making them work together. But what I am saying is that figuring out where the true insights are and finding the way of making them work together is the work of fundamental physics. This work is grounded on the confidence in the old theories, not on random search of new ones. … The ‘wild’ scientist observes that great scientists had the courage of breaking with old and respected ideas and assumptions, and explore new and strange hypothesis. From this observation, the ‘wild’ scientist concludes that to do great science one has to explore strange hypotheses, and violate respected ideas. The wildest the hypothesis, the best. I think wilderness in physics is sterile. The greatest revolutionaries in science were extremely, almost obsessively, conservative. Now, we are in full agreement with this “conservative” view on fundamental physics, against the popular “wilderness”. Based on this common methodological background, we disagree mainly in one point: in the decisions what are the “true insights” of the old theories which should be taken seriously and extended into the domain of quantum gravity, and which should be explained, derived, and therefore not extended into quantum gravity. Now, based on these methodological rules, we present the existence of a preferred frame as the deep insight of Bohmian mechanics. In other parts we criticize the usual “insights” of general relativity: the relativistic notion of “time” (§ C) and relationalism (§ B). ### 14.1 The insights of Bohmian mechanics In some sense, the disagreement starts with the definition of the theories we want to unify in quantum gravity. Usually this is presented as the problem of unification of general relativity and special-relativistic quantum field theory. But it is completely ignored that there exists already a quantum theory of gravity – multi-particle Schrödinger theory for the Newtonian interaction potential. And, moreover, there is an interesting variant of this theory – Bohmian mechanics (BM). While it is in no way suggested that the interesting and important insights of quantum field theory should be ignored, the theories we really have to unify are general relativity and non-relativistic Bohmian mechanics. We consider some details of BM in appendix F. This theory proves that vagueness, subjectivity, and indeterminism of the usual interpretations of quantum theory are not forced on us by the experimental facts. These are very interesting and important insights into the nature of quantum theory – in my opinion, insights of much more fundamental character in comparison with the domain of applicability of particular space-time symmetries. The problem with BM is that it requires a preferred frame, in contradiction with the relativity principle. In this context, it seems useful to quote again Rovelli : “So, Einstein believed the two theories, Maxwell and Galileo. He assumed that they would hold far beyond the regime in which they had been tested. … If there was contradiction in putting the two together, the problem was ours: we were surreptitiously sneaking some incorrect assumption into our deductions.” Now, is there a possibility to believe above theories? If we require Lorentz invariance on the fundamental level, Bohmian mechanics should be simply rejected and does not give any insight. This certainly violates the recommendation to take above theories seriously. The choice of GET is to preserve Lorentz invariance for observable effects, but to accept a preferred frame on the fundamental level. Now, the question is how much this weakens the insights of relativity. Obviously, we do not destroy the insights of relativity completely. Instead of “the stage does not exist” we obtain “the stage is not observable”. This remains to be an important and non-trivial insight. ### 14.2 Relativity as a theory about observables But we want to go farther. We argue that relativity is only a theory about observables. Therefore, nothing changes if we replace “the stage does not exist” with “the stage is not observable”. Indeed, if there would be a difference, then there should be a method to show the existence of unobservable objects. In classical realism, there are such methods – the EPR criterion of reality allows to prove the existence of such hidden objects. But this criterion has been rejected. Moreover, this is a necessity. Without the rejection of the EPR criterion relativity would be simply falsified. Thus, after the rejection of the EPR criterion we have no longer a chance to prove the existence of an unobservable stage. Lorentz invariance is only observable Lorentz invariance. The original principle “all laws of nature should be Lorentz-invariant” has been replaced by “all observable effects should be Lorentz-invariant” at least after Aspect’s experiment. Therefore, from point of view of relativity the statements “the stage does not exist” and “the stage is not observable” are simply identical. If we interpret the relativistic insight “the stage does not exist” as “the stage is not observable” we do not weaken relativity. Instead, we take the relativistic preference for observables seriously. Thus, we conclude that the acceptance of a preferred frame is in natural agreement with the methodology recommended by Rovelli, which requires to take above theories seriously. ## 15 The violation of Bell’s inequality Before learning the details of the violation of Bell’s inequality, I have thought that it gives some weak evidence in favour of a preferred frame. It was a real surprise for me that the evidence is, instead, very strong – its not an exaggeration to name it simply a falsification of relativity. The whole problem is a very simple one. While relativity forbids any superluminal causal influence, the Lorentz ether allows them, as long as their observable consequences allow two explanations: ($`AB`$ or $`BA`$). In this case, they cannot be used to measure absolute time, which is forbidden in the Lorentz ether. Now, Bell’s inequality may be violated only if ($`AB`$ or $`BA`$). And, because it is violated, relativity is falsified and we have to return to the Lorentz ether. A very nice, interesting but simple example of an indirect existence proof. But, instead of accepting this elementary experimental falsification, the simple but fundamental principles used in this proof are questioned, even rejected. The argumentation used in this context is a classical example of immunization. Some arguments are nonsensical enough to describe the situation as a “flight from reason in science” (Goldstein ). Fortunately, reading Kuhn’s “structure of scientific revolutions” has recovered my optimism about the presence of reason in science. Kuhn observes that paradigms are never falsified by experiment alone. Their rejection always requires another paradigm for replacement. And, without GET, the ether paradigm was simply not a viable competitor. But with GET the ether paradigm becomes a reasonable competitor of relativity. Now, with GET as the background, the violation of Bell’s inequality should be reinterpreted. From point of view of competition between ether and relativity it becomes a simple and beautiful experimental falsification of relativity. Of course, there seems no necessity to copy the well-known proofs of the various variants of Bell’s inequality. Nonetheless, to explain some features (like the simplicity of the theorem itself, the classical character of the decisions and the observations and the existence of applications) the simplest way seems to be a simple “proof for schoolboys”: ### 15.1 Bell’s inequality for schoolboys There are three cards, left, middle, right, with red or black color. I have to choose them so that: * the left and middle cards have the same color; * the middle and right cards have the same color; * the left and right cards have different color. Obviously, one of the three claims is wrong. Now, if you open two cards, you can test one of these claims. What’s the probability p to detect a wrong claim? Obviously $`p1/3`$. What if you win only with $`p<1/3`$? Obviously, something is manipulated. What? That’s easy: after you have chosen the first card, another card may be manipulated. For example, imagine one card is marked. If you open this card, you hand becomes marked. And the other card may be manipulated so that if touched by a marked hand it changes its color. Let’s try to avoid this possibility for manipulation. We use two rooms and assume that no information transfer between the two rooms is possible. In every room we have three cards, and your team has to choose one card in every room. How can you be sure that the cards in the two rooms are the same? Very simple, you can ask for the same card in above rooms as often as you like, and in this case these cards should always have the same color. Without information about the question in the other room your opponents cannot be sure if you ask about different cards or the same cards. They have no better strategy than to use the same predefined color for the same card. Therefore, we are in the same situation as before, but without the possibility to manipulate based on the information about the other question. Therefore, we have again $`p1/3`$. And, if not, there is a hidden information transfer from one room to the other. That’s all – we have proven Bell’s theorem. Let’s formulate it in the following way: ###### Theorem 4 (Bell’s theorem) If Bell’s inequality $`p1/3`$ is violated for measurements at A and B, then there exists a causal influence ($`AB`$ or $`BA`$). ### 15.2 The difference between special relativity and Lorentz ether As far, we have used only classical common sense, and proven that from the violation of Bell’s inequality follows ($`AB`$ or $`BA`$). The question is now how relativity and ether theory are involved. In special relativity, we have no absolute time. Instead, in pre-relativistic Lorentz-Poincare ether theory we have an absolute time, but we cannot measure it. Because of the ether, moving clocks are dilated and moving rulers contracted. Now, the formulas for SR and Lorentz ether are identical. Therefore, it is often said that above theories are identical in their predictions. But this is not true. The violation of Bell’s inequality is an interesting difference. In special relativity we have Einstein causality. This is a consequence of two axioms: that there are no causal loops, and that causality is a law of nature, and therefore has to be Lorentz invariant. Einstein causality is Lorentz-invariant, but a notion of causality which allows faster than light causal influences leads to causal loops. Therefore, any causal influence faster than light is forbidden. That means, as $`AB`$, as $`BA`$ is forbidden by Einstein causality. Therefore, ($`AB`$ or $`BA`$) is forbidden too. But in the Lorentz ether the situation is different. We have classical causality. Therefore, causal influences faster than light are not forbidden. There is only one restriction: all observable effects should be Lorentz-invariant. Now, it seems to follow that such faster than light influences cannot have observable consequences. But that’s wrong – observable effects between space-like separated events A and B may be very well Lorentz-invariant if they allow two explanations: ($`AB`$ or $`BA`$). If in absolute time $`t_A<t_B`$, we use the explanation $`AB`$, but if $`t_A>t_B`$, we use explanation $`BA`$. Thus, ($`AB`$ or $`BA`$) is not forbidden in the Lorentz ether. We conclude that there is an interesting difference in the predictions of special relativity and Lorentz ether theory: observable correlations between space-like separated events A and B which may be explained by causal influences ($`AB`$ or $`BA`$) are forbidden in special relativity, but allowed in Lorentz ether theory. Bell’s inequality is a simple example of an effect of this type. We have above explanations. If Alice is able to send information to Bob, they can always win. If Bob is able to send information to Bob, they can always win too. Thus, if they win with probability $`p<1/3`$, we can explain this as ($`AB`$ or $`BA`$). Violations of Bell’s inequality are obviously observable. We see that it is not correct to claim that special relativity and Lorentz ether are equivalent as physical theories. They are not. In special relativity, we can prove Bell’s inequality for space-like separated events. In the Lorentz ether violations of Bell’s inequality are allowed. ### 15.3 Aspect’s device Now, there is no need to understand how Aspect’s device works. It’s sufficient to consider it as a black box, or, more accurate, a device consisting of two black boxes, one for each room. You can press one of three buttons – left, middle, right – and it gives the answer – red or black. And if Alice and Bob use this device, your probability to find the wrong answer is $`p=1/4`$. That’s all. We conclude that we have to reject special relativity and to return to the Lorentz ether. Bell’s conclusion was similar: “the cheapest resolution is something like going back to relativity as it was before Einstein, when people like Lorentz and Poincare thought that there was an ether — a preferred frame of reference — but that our measuring instruments were distorted by motion in such a way that we could no detect motion through the ether. Now, in that way you can imagine that there is a preferred frame of reference, and in this preferred frame of reference things go faster than light.” ### 15.4 Should we question everything? We have presented a very simple proof of Bell’s inequality. Unfortunately proofs are certain only in mathematics. As far as we consider reality, nothing is certain. And, therefore, everything in this proof may be (and has been) questioned. This reaction is justified in a situation where relativity is an unquestioned paradigm of science. But it is no longer justified in a situation where we have a competition between two paradigms – relativity vs. ether theory. In this situation, the previously justified search for explanations becomes a simple method of destruction – a method which may be applied to every experimental falsification of a physical theory: take an arbitrary sentence in the proof we have presented. Then, name it an “unproven hypothesis”, explain that we “cannot be sure” that it holds, and that “far away from our everyday experience” this common sense assumption is possibly invalid. That’s sufficient, you have found a “loophole” in the proof. Moreover, you will have a lot of fun if somebody tries to justify the statement you have questioned – especially because he cannot be successful. Whatever he presents, you have a simple counter-attack: take an arbitrary sentence from his reply, and … you already know. This is the method described by Rovelli as “wilderness”: > The ‘wild’ scientist observes that great scientists had the courage of breaking with old and respected ideas and assumptions, and explore new and strange hypothesis. From this observation, the ‘wild’ scientist concludes that to do great science one has to explore strange hypotheses, and violate respected ideas. The wildest the hypothesis, the best. I think wilderness in physics is sterile. The greatest revolutionaries in science were extremely, almost obsessively, conservative. That we have to stop somewhere with the questioning is well-known in scientific methodology. Every test of a theory requires to stop questioning at some point. That’s a well-known fact in scientific methodology. For example, Popper writes: > Every test or theory … must stop at some basic statement or other which we decide to accept. If we do not come to any decision, … then the test will have led nowhere…. Thus if the test is to lead us anywhere, nothing remains but to stop at some point or other and say that we are satisfied, for the time being. Note that this is not dangerous at all: we do not have to accept something forever, in a dogmatical way, we can remove our acceptance whenever we have reason to doubt. Now, the question may be only what are reasonable criteria to stop questioning some principles, at least for time. In our case, a simple rule seems to be that an objection should be justified by something more than the simple and trivial remark that it is an unproven hypothesis. Based on this basic rule let’s reject now some well known objections against our resolution of the violation of Bell’s inequality. Thus, unjustified, pure doubt into some common sense principle, without reason, that means without independent evidence against the questioned principle, should be rejected. Let’s look now at the common objections against the obvious resolution of the violation of Bell’s inequality – the acceptance of a preferred frame. ### 15.5 No application for the violation One common counter-argument is that we cannot use the violation of Bell’s inequality for information transfer. This is really strange, because to use it for information transfer is forbidden in the Lorentz ether too. It is simply in contradiction with the requirement that two explanations of the violation are possible: ($`AB`$ or $`BA`$). If we use this device for information transfer $`AB`$ this is certainly in contradiction with the explanation $`BA`$. Therefore, we have the strange situation that it is considered to be an argument against a theory that the prediction of this theory holds. A variant of this argument is that the violation of Bell’s inequality cannot be applied. That’s simply wrong. In our proof, we have described an application – Aspect’s device allows Alice and Bob to improve their coordination and to win in the game with probability $`p=3/4`$ instead of $`2/3`$. But to improve coordination in similar game-like situations is the classical application of information transfer. Therefore, even if pure communication is impossible – for known reasons – we can apply it in a similar way to improve coordination. Moreover, it should be noted that this objection, while common, is in no way an argument in favour of one of the competitors. ### 15.6 Objections against a preferred frame One argument which was really justified in the past is that there was no theory of gravity with a preferred frame comparable with general relativity. Now, with the theory presented here this argument is obsolete. Of course, all other known arguments against the Lorentz ether may be presented in this context. Most of them, again, are obsolete in GET. One specific argument is worth to be considered: It is argued that it is not natural to use two different explanations for the same observation. But two Lorentz-equivalent configurations are not “the same” in the Lorentz ether. That’s the general situation in the Lorentz ether. In different but Lorentz-equivalent situations all other observable effects are explained in a different way. In this context it is natural to have different explanations for the violation of Bell’s inequality too, much more natural than to have the same explanation. Now, you may like or not like the properties of the Lorentz ether – the question is if ($`AB`$ or $`BA`$) is allowed in the Lorentz ether or not, if there are arguments against the experimental falsification of relativity. To count here metaphysical arguments against the other competitor is really strange. If we accept arguments about metaphysical beauty as decisive against an experimental falsification, this is simply the end of science. ### 15.7 No contradiction with quantum mechanics Then where is a whole class of common objections: objections related with the strangeness of quantum theory. The strangeness of the double slit experiment or the quantum eraser seems to suggests that the preferred frame is not the problem, and that the common sense principles we have used in our proof are in contradiction with quantum theory. Of course, we can argue here that the burden of proof is on the side of the people who claim that there is such a contradiction. But we don’t have to – instead, we can prove that there is no such contradiction. ###### Theorem 5 There is no contradiction between the principles used in the proof of Bell’s theorem and the predictions of non-relativistic quantum theory. The proof is given by Bohmian mechanics – a hidden variable theory of quantum mechanics found by David Bohm . In a “quantum equilibrium”, this theory makes the same predictions as non-relativistic quantum theory, therefore, no prediction of non-relativistic quantum theory is in contradiction with the predictions of Bohmian mechanics. But Bohmian mechanics is in full agreement with the common sense principles used in the proof of Bell’s theorem. We consider Bohmian mechanics in appendix F. Thus, “quantum strangeness” is not a reason to reject the proof. All arguments related with double slit experiments, spin, quantum erasers, the “wholeness” of quantum theory and so on are irrelevant. It gives no support for the thesis that something is wrong with the proof of Bell’s theorem. ### 15.8 Objections in conflict with Einstein causality There is another class of objections which may be rejected as a whole. It is based on the assumption that Einstein causality is considered as a prediction of relativity. The idea is a detailed consideration of an assumed faster than light phone line. Let’s assume such a phone line exists. In this case, we obviously have to give up relativity as falsified. Nonetheless, let’s consider the seemingly nonsensical question how we can prove this based on the observation. The point is that we can play here the same game of “questioning everything”. Now, in this situation this looks certainly nonsensical. Surprisingly, we look at the details, we find that we have to use assumptions which may be and have been questioned in the discussion about Bell’s inequality. We propose the following criterion: all assumptions and principles which have to be used to prove that a really working superluminal phone line falsifies Einstein causality should not be questioned in the proof of Bell’s theorem. Now, let’s look how to prove this. Let’s consider the basic device which transfers one bit of information. The FTL phone consists of two black boxes. We make some decision and press a button on our box. Our friend makes a measurement on his side of the box and obtains a result. Then, we meet later and compare our decisions with his observations. If we observe a significant correlation we conclude that the phone works. We see that the experimental situation is very close to the situation in the violation of Bell’s inequality, and really includes parts which have been questioned: * The device is a black box, we don’t know how it works. This is similar to our presentation of Aspect’s device as a black box. Therefore, every argumentation which requires some insight into this black box before accepting that some causal information transfer has happened should be rejected. * We essentially use the free will of the experimenter. Of course, as in Bell’s theorem, the experimenter may be replaced by various other random number generators to obtain certainty that the input – the decision of the experimenter – is not predefined. But that’s enough. Every argument which does not accept this as sufficient to establish the independence of the decision of the experimenter should be rejected. * We can verify the existence of a non-trivial correlation only after the experiment has finished, using other methods of information transfer. It is typical for “many world” explanations or explanations based on “superpositions of observers”: nothing non-local happens, only if the observers meet again, with usual subluminal speed, something collapses and we obtain the correlations in a miraculous quantum way. These explanations should be rejected too. * Note also the general objection what we observe only a correlation. Having only a correlation, we cannot conclude that there really exists a causal relation in reality. We need some principle which allows to make the step from observable correlations to the conclusion that there exists a real causal interaction. Every criticism which rejects the way this has been done in the proof of Bell’s inequality but does not describe an alternative way to conclude from observation that a real, causal relation exists should be rejected. * We know that as $`AB`$, as $`BA`$ is forbidden by Einstein causality. Therefore, ($`AB`$ or $`BA`$) is forbidden too. This is obvious. But there exists a well-known psychological bias known as “disjunction fallacy” : if there are two alternatives, and above alternatives lead to the same conclusion (in our case: the falsification of relativity), but we are not sure which alternative happens, we tend not to make the conclusion. It seems, this fallacy is the base for arguments of type “we are unable to detect the direction of influence”. Again, our argumentation allows to reject a whole class of common objections. Of course, it works only if we want to defend Einstein causality as a physical prediction of relativity. It does not work against the idea of rejection of causality itself, for example following Price . ### 15.9 Discussion Bell’s inequality is a prediction of SR which does not hold in LET. Therefore the violation of Bell’s inequality falsifies SR and requires to accept a preferred frame. The proof is indirect but so simple and straightforward that it is hard to imagine a stronger indirect falsification of a physical theory. The consideration of the common counter-arguments has not given any serious loopholes. Instead, most of the common arguments are part of two classes we have rejected: objections based on assumed contradictions with quantum principles, and objections which, if accepted, would allow to immunize relativity even if we have working superluminal phone lines. What remains is the possibility to question everything. Certain proofs exist only in mathematics, not in physics. Immunization of a physical theory is always possible. In principle, objections are always possible. Last not least, we are talking about reality, not about pure mathematics. But there is no independent evidence which suggests that some part of this proof should be questioned. Therefore, the rejection of the proof of Bell’s inequality has to be qualified as an ad hoc immunization of relativity. Nonetheless, in § E we consider the EPR criterion in more detail. We present there evidence for the thesis that the principles used in the proof of Bell’s inequality – principles we denote as “EPR realism” – are of more fundamental character than space-time symmetries. ## 16 Conclusions General ether theory proposes a paradigm shift back from relativity to a classical Newtonian background. It heals the main problems of the old Lorentz ether: * relativistic symmetry is explained in a general, simple way; * the ether is generalized to gravity; * the ether is compressible, changes in time; * there is no longer an action-reaction problem; In various parts of fundamental physics we find advantages of the new approach: * we obtain local energy and momentum density for gravity; * the problem of time of quantum gravity is solved; * singularities in physical important situations disappear; * frozen stars instead of black holes solve the information loss problem of quantum gravity; * we obtain a reasonable dark matter term; * a big bounce instead of a big bang solves the cosmological horizon problem without inflation; * the EPR criterion of reality holds; * a generalization of Bohmian mechanics into the relativistic domain becomes unproblematic; Even in the domain of beauty GET seems to be able to compete with GR: it seems that some of the most beautiful aspects of the mathematical apparatus of GR obtain a physical interpretation in the context of GET: harmonic coordinates, ADM decomposition, triad formalism, Regge calculus. Especially the beautiful relation between covariance and conservation laws has to be mentioned here: it gives in GET two nice expressions for the conservation laws, but in GR it makes energy conservation highly problematic. Reconsidering the metaphysical foundations of relativity, we have found serious flaws: * A reconsideration of the violation of Bell’s inequality suggests its interpretation as a falsification of relativity; the rejection of the EPR criterion of reality should be interpreted as an immunization of relativity: except its contradiction with the relativity principle there is no independent evidence against the EPR criterion. * A new quantization problem – named scalar product problem – suggests that a really relativistic covariant quantum theory fails to describe the non-relativistic limit. * The consideration of the “insight” of relativity into the nature of time suggests that it is based on conflation of different physical notions – clock time and “true” time. All they have in common is that they have been named “time”. * The consideration of general covariance and relationalism shows that these are not advantages of relativity, but can always be reached – by forgetting valuable information. On the other hand, we have been unable to detect serious difficulties of the new approach: there is no evidence that the introduction of a preferred frame is problematic in any part of modern physics. But, of course, a lot of interesting questions remain open, especially the physical interpretation of gauge fields and fermions, Lagrange formalisms for condensed matter compatible with the GET scheme, the details of the related hidden relativistic symmetry in usual condensed matter. The unification of the geometrical methods developed for GR quantization with the canonical quantization schemes for condensed matter will be interesting for above domains: it defines how to regularize in quantum gravity, and it gives geometrical interpretation in condensed matter theory. ## Appendix A Covariant description for theories with preferred coordinates The covariant description we consider here is quite simple: we handle the preferred coordinates formally as “scalar fields” $`X^\mu (x)`$. This allows to describe a theory with preferred coordinates in a covariant way. An interesting point of this covariant description is that the Euler-Lagrange equations for the preferred coordinates are the conservation laws. This is quite obvious and may be considered as a “folk theorem”. Nonetheless, confusion about the physical meaning of “covariance” seems quite common, and it does not seem to be widely known. That’s why it seems reasonable to describe the “weak covariant” description we use in more detail. ### A.1 Making the Lagrangian covariant We assume that we live in a Newtonian framework. That means, there is an absolute Euclidean space and absolute time. To describe this absolute background, we use preferred coordinates $`X^i,T=X^0`$. As usual, Latin indices vary from 1 to 3, Greek indices from 0 to 3. Of course, to use other coordinates $`x^\mu `$ is not forbidden, the coordinates $`X^i,T`$ are only preferred – the laws of nature are simpler in these coordinates. Now let’s consider how to obtain a covariant description starting from a non-covariant one: ###### Theorem 6 Let $`S=L(T_{\mathrm{}}^{\mathrm{}},_\mu T_{\mathrm{}}^{\mathrm{}})`$ be a functional which depend on components $`T_{\mathrm{}}^{\mathrm{}}`$ and first derivatives $`_\mu T_{\mathrm{}}^{\mathrm{}}`$ of tensor fields T. Then there exists a covariant functional $`S_c=L_c(T_{\mathrm{}}^{\mathrm{}},_\mu T_{\mathrm{}}^{\mathrm{}},X_{,\nu }^\mu )`$ which depends on the components and first derivatives of the same tensor fields T and on the first derivatives $`X_{,\nu }^\mu `$ of four scalar fields $`X^\mu (x)`$ so that $$S(T)=S_c(T,X_{,\nu }^\mu )$$ for $`X^\mu (x)=x^\mu `$. Proof: In L we replace the tensor components by expressions using the following replacement rules for all indices: $`T_{\mathrm{}}^\mathrm{}\mu \mathrm{}`$ $``$ $`{\displaystyle \frac{X^\mu }{x^\nu }}T_{\mathrm{}}^\mathrm{}\nu \mathrm{}`$ $`T_\mathrm{}\mu \mathrm{}^{\mathrm{}}`$ $``$ $`{\displaystyle \frac{x^\nu }{X^\mu }}T_\mathrm{}\nu \mathrm{}^{\mathrm{}}`$ $`T_{\mathrm{},\mu }^{\mathrm{}}`$ $``$ $`{\displaystyle \frac{x^\nu }{X^\mu }}T_{\mathrm{},\nu }^{\mathrm{}}`$ $`T_\mathrm{},^\mathrm{}\mu \mathrm{}`$ $``$ $`{\displaystyle \frac{X^\mu }{x^\nu }}T_\mathrm{},^\mathrm{}\nu \mathrm{}`$ $`T_{\mathrm{}\mu \mathrm{},}^{\mathrm{}}`$ $``$ $`{\displaystyle \frac{x^\nu }{X^\mu }}T_{\mathrm{}\nu \mathrm{},}^{\mathrm{}}`$ The matrix $`\frac{x^\nu }{X^\mu }`$ is the inverse matrix of $`\frac{X^\mu }{x^\nu }`$. We use this property to express all occurrences of $`\frac{x^\nu }{X^\mu }`$ by these rational functions of $`X_{,\nu }^\mu `$. For $`X^\mu (x)=x^\mu `$ these are obviously identical transformations. Moreover, each argument is now a covariant scalar. Indeed, the indices $`\mu `$ of the $`X_{,\nu }^\mu `$ are not tensor indices, but enumerate the scalar fields $`X^\mu (x)`$. But indices of this type are the only open indices in the expression. Therefore, being a function of covariant scalar expressions, the modified function $`L_c`$ is a covariant scalar function, qed. Covariant functions like $`L_c`$ which do not depend on the fields $`X^\mu (x)`$ we name “strong covariant”, to distinguish them from “weak covariant” functions which depend on the fields $`X^\mu (x)`$. Of course, what is not part of this theorem is how to interpret something defined only in the preferred coordinates as a component of some tensor field. This is, essentially, the real problem if we have to make a theory covariant. There are usually different possibilities: a 3D scalar may be a 4D scalar as well as a component of a 4D vector or tensor field. But this is a question of the definition of the theory itself. We are interested here only in a special way to obtain a covariant description of a well-defined theory. ### A.2 Conservation laws in the covariant formalism This formalism raises two interesting questions. First, once we have a covariant formalism, we obtain the known problems with conservation laws – Noether’s theorem does not give non-trivial conservation laws. Second, we have four new Euler-Lagrange equations for $`S=L`$ – the Euler-Lagrange equations for the preferred coordinates $`X^\mu `$. What is their physical meaning? Now, the answer is simple and beautiful – the Euler-Lagrange equations for the preferred coordinates are the conservation laws. All we need is to look at the Euler-Lagrange equations for the preferred coordinates: ###### Theorem 7 If a Lagrangian L does not depend explicitly on the coordinates, then the Euler-Lagrange equation of the related weak covariant Lagrangian $`L_c`$ for the preferred coordinates $`X^\mu `$ defines conservation laws for the tensor $$T_\mu ^\nu =\frac{L_c}{X_{,\nu }^\mu }$$ If the original Lagrangian L is covariant, then $`T_\mu ^\nu =0`$. If there was no explicit coordinate-dependence in the original Lagrangian, there is none in $`L_c`$ too. Note also that for a covariant Lagrangian $`L`$ we have $`LL_c`$ and does not depend on $`X^\mu `$ and $`X_{,nu}^\mu `$. This follows from the construction. The theorem now follows immediately from the the Euler-Lagrange equation for the $`X^\mu `$: $$\frac{\delta S}{\delta X^\mu }=\frac{L}{X^\mu }_\nu \frac{L}{X_{,\nu }^\mu }=0$$ The results about the existence of conservation laws are equivalent to Noether’s theorem, but the energy-momentum tensor is not the same. The relation between preferred coordinates and conservation laws is a much more direct one in this formalism: de facto one line was sufficient for the proof. We have not used the other Euler-Lagrange equations. That’s why we consider this variant of the conservation law as the more fundamental one. We conclude: the Euler-Lagrange equations for the preferred coordinates are the conservation laws. ## Appendix B Relationalism The confusion about covariance has historical reasons. Initially covariance was believed to be a special, distinguishing property of GR. Later it has been recognized that every physical theory allows a covariant formulation. For example, Fock has given a covariant formulation for special relativity. A simple way to do this is to use the curvature tensor of a metric $`g_{ij}`$. The equation $`R_{jkl}^i=0`$ defines a flat metric in a covariant way. Once a flat background has been defined, all partial derivatives of the equations in the preferred frame may be replaced by covariant derivatives of this background metric. This method has been widely used to present theories of gravity with predefined geometries, for example , . Nonetheless, this does not mean that this question is completely clear now. The situation is confusing. On one hand, every theory allows a covariant formulation, on the other hand, there is a non-trivial symmetry property – the property we have named “strong covariance” in GET. Rovelli describes it using the notions “active vs. passive diff invariance”: > All this is coded in the active diffeomorphism invariance (diff invariance) of GR. Passive diff invariance is a property of a formulation of a dynamical theory, while active diff invariance is a property of the dynamical theory itself. A field theory is formulated in manner invariant under passive diffs (or change of coordinates), if we can change the coordinates of the manifold, re-express all the geometric quantities (dynamical and non-dynamical) in the new coordinates, and the form of the equations of motion does not change. A theory is invariant under active diffs, when a smooth displacement of the dynamical fields (the dynamical fields alone) over the manifold, sends solutions of the equations of motion into solutions of the equations of motion. This is in agreement with our understanding. The problematic part is how to distinguish dynamical and non-dynamical fields. Rovelli continues: > Distinguishing a truly dynamical field, namely a field with independent degrees of freedom, from a non-dynamical field disguised as dynamical (such as a metric field $`g`$ with the equations of motion Riemann\[g\]=0) might require a detailed analysis (for instance, Hamiltonian) of the theory. That means, it is assumed that an analysis of the equations of the theory – the dynamics – allows to distinguish the non-dynamical and the dynamical steps of freedom of the theory. The example of GET and its relation with general relativity with four scalar dark matter fields (GRDM) allows to proof that this is impossible: Thesis: It is impossible to distinguish a non-covariant theory with a “preferred frame in disguise” from a “truly” covariant physical theory by evaluation of the equations of motion and the Lagrange formalism. Indeed, in GET, the preferred coordinates $`X^\mu `$ are non-dynamical, and their presentation as “dynamical fields” $`X^\mu (x)`$ fits exactly with the description of a “non-dynamical field disguised as dynamical”. As we have seen, the equation for the preferred coordinates $`X^\mu (x)`$ is simply the harmonic equation, thus, the usual general-relativistic equation for scalar fields. On the other hand, in GRDM the four scalar fields are dynamical. But the Lagrange formalism for GET is exactly the Lagrange formalism for GRDM. A variant of this argument is the consideration of GET with the four “preferred coordinates in disguise” $`X^\mu (x)`$ and a few additional free scalar fields $`\phi ^m(x)`$. Then, looking at the equations and the Lagrangian, there is no way to tell what are the truly dynamical scalar fields $`\phi ^m(x)`$ and what are the coordinates in disguise $`X^\mu (x)`$. We need additional a-priori information, additional insight. Once the equations are identical, even a “Hamiltonian analysis” (whatever this means in detail) cannot help. But there is an interesting point: the Hamilton formalism is different for GET and GRDM. In GRDM we have the typical problems of general-relativistic theories where the Hamiltonian is a constraint. Instead, in GET we have a classical Hamilton formalism, and the Hamiltonian is not a constraint. But we cannot derive the Hamilton formalism of GET without the additional information that the fields $`X^\mu (x)`$ are the preferred coordinates. Without this information we cannot decide which Hamilton formalism is the appropriate one. Note that our thesis does not mean that there are no differences between GET and the “truly relativistic” theory GRDM. The differences we have considered in § 7. Our thesis is that we need additional insight – the insight that the fields $`X^\mu `$ are preferred coordinates – to be able to distinguish a theory with absolutes from a “truly relativistic” theory. ### B.1 What is the true “insight” of general relativity? This observation seems to destroy the whole concept described by Rovelli : > One of the thesis of this essay, is that general relativity is the expression of one of these insights, which will stay with us “forever”. The insight is that the physical world does not have a stage, that localization and motion are relational only, that diff-invariance (or something physically analogous) is required for any fundamental description of our world …. > > In GR, the objects of which the world is made do not live over a stage and do not live on spacetime: they live, so to say, over each other’s shoulders. But, as we have seen, this is not an insight, but a tautological reformulation. If the theory has a stage, we can reformulate it and name the stage a dynamical field. Instead of “absolute motion” we talk about “motion relative to the stage field”. Let’s see how Rovelli describes Einstein’s “insight”: > Of course, nothing prevents us, if we wish to do so, from singling out the gravitational field as “the more equal among equals”, and declaring that location is absolute in GR, because it can be defined with respect to it. But this can be done within any relationalism: we can always single out a set of objects, and declare them as not-moving by definition. The problem with this attitude is that it fully misses the great Einsteinian insight: that Newtonian spacetime is just one field among the others. The situation between GET and GRDM is the reverse one. The fields $`X^\mu (x)`$ are not just scalar fields among others. But this is a non-trivial insight. If we forget this insight, and the fields $`X^\mu (x)`$ are interpreted just as fields among others, we have lost important information and interesting predictions. In this way, by forgetting interesting information, a relational description is always possible. Therefore, the existence of a relational description is only an insight into the mathematical formalism of physical theories in general, not an insight of GR. Instead, in GET the existence of absolutes – the preferred coordinates $`X^\mu (x)`$ – is a non-trivial insight which leads to interesting predictions: the fields $`X^\mu (x)`$ may be used as global coordinates, the field $`X^0(x)`$ is time-like. ### B.2 What are insights which are “forever”? What would be an important insight is that a theory with a certain type of absolutes is impossible in nature. But for this insight general relativity is not enough. This can be, by its nature, only an impossibility proof for certain classes of theories. Such an insight may be based on certain observational facts, for example, observations of a worm-hole. This would be incompatible with a whole class of theories with flat background. Such an observation of a non-trivial topology, for example a worm-hole, would be really an insight which remains forever.<sup>11</sup><sup>11</sup>11Another example of this type of insight is the violation of Bell’s inequality, which excludes a whole class of theories – theories which allow the proof of Bell’s inequality for space-like separated observations. This can be named also an “insight of quantum theory”. But this would be a very sloppy description: the proof of Bell’s is classical, and the violation of Bell’s inequality has been observed in real experiments. But we have not observed such non-trivial topology. A lecture which can be learned from history of science is that metaphysical preferences like between relationalism and the existence of a predefined stage should not be justified based on the current physical theory. Rovelli himself has presented the best example of this type – Newton’s insight, related with the famous rotating bucket, that there exists an absolute space. This insight, which was thought to remain forever too, was rejected by Einstein. In Einstein’s theory, we have relationalism, no absolutes. But this insight may be as well superseded by another theory. GET proves that this is possible. There is another example of an interesting metaphysical question where we tend to use current physical theory: probability versus determinism. Here we have even more switches between probabilistic chaos and determinism. Bohmian theory (deterministic), quantum theory (probabilistic), classical mechanics (deterministic), chaos in the many-particle situation (probabilistic), classical thermodynamics (deterministic), and chaos in its large scale predictions. Therefore, to base the metaphysical decision between chaos and determinism on current physical theory is highly speculative. To guess that this property remains forever is sufficiently falsified by historical evidence. There are important insights of general relativity. For example that a special physical entity which was absolute in Newton’s theory is not absolute in reality. It is the entity which defines inertial forces and causes the difference between a rotating bucket and a bucket in rest. This entity is the gravitational field. This insight into the nature of gravity and clock time will stay forever. Another insight is the Einstein equivalence principle. Because this is an exact symmetry claim, we cannot be sure that it remains forever – but we can be sure that it survives at least as an approximative symmetry. ## Appendix C The problem of time It is well-known that the problem of time may be solved by the introduction of a preferred foliation as in GET: “in quantum gravity, one response to the problem of time is to ‘blame’ it on general relativity’s allowing arbitrary foliations of spacetime; and then to postulate a preferred frame of spacetime with respect to which quantum theory should be written. Most general relativists feel this response is too radical to countenance: they regard foliation-independence as an undeniable insight of relativity.” ### C.1 clock time vs. true time To meet this argument, we have to consider the “insight of relativity” about the nature of time and the metaphysical aspects of the “problem of time” in more detail. In some sense, the problem of time may be discussed based on Newton’s definition . Newton distinguishes two notions of time: > … I do not define time, space, place, and motion as being well known to all. Only I observe, that the common people conceive these quantities under no other notions but from the relation they bear to sensible objects. And thence arise certain prejudices, for the removing of which it will be convenient to distinguish them into absolute and relative, true and apparent, mathematical and common. > > I. Absolute, true and mathematical time, of itself, and from its own nature, flows equable without relation to anything external, and by another name is called duration: relative, apparent, and common time, is some sensible and external (whether accurate or unequable) measure of duration by means of motion, which is commonly used instead of true time; such as an hour, a day, a month, a year. This definition is in agreement with GET. Here, we have an unobservable harmonic “true time” $`T(x)`$ together with the “apparent time” $`\tau =\sqrt{g_{\mu \nu }dx^\mu dx^\nu }`$ measured with clocks. These two notions of time may be roughly identified with “time of quantum theory” and “time of relativity”: Indeed, in quantum theory we have no self-adjoint operator for time measurement. Instead, it is closely related with the fundamental aspects of quantum theory: physical quantities have to be measured at a given time, the scalar product is conserved under time evolution. Thus, it ideally fits with Newtons “true time” and should be distinguished from “apparent time”. On the other hand, time in relativity is defined to be apparent time. General-relativistic time is the time measured by clocks – “measure of duration by means of motion”. And its most interesting physical property is that it is unequable. The fit with Newton’s definition of apparent time is simply ideal. Therefore, if we deny Newton’s insight into the different nature of true and apparent time, and try to develop quantum gravity without making this distinction, we are immediately faced with the problem to unify these two different notions of time. This is the metaphysical base of the notorious “problem of time” in quantum gravity. ### C.2 About positivistic arguments against true time The standard relativistic argumentation against this understanding of time is the rejection of “true time” based on positivistic arguments. The typical argument is that a physical notion needs an operational definition. If we cannot measure something, then it is not part of physics. Now, this argumentation is based on positivistic methodology of science, which has been rejected by Popper . According to Popper’s “logic of scientific discovery” theory is prior to observation. Therefore, the fundamental objects of a theory are not based on observation, they do not need any operational definition. It is not the observation which decides about the fundamental notions of the theory. Instead, observation is always theory-laden. In Einstein’s famous words, it is the “theory which decides what is observable”. Positivism reverts the relation between theory and observation. All what is required in science is that the theory, as a whole, makes a lot of falsifiable predictions. An operational definition of some fundamental notions of the theory is a useful tool to obtain such predictions, but not more. Positivism is simply wrong. Again, in Einstein’s words about Bohr’s positivism: ”Perhaps I did use such a philosophy earlier, and even wrote it, but it is nonsense all the same.” Moreover, in our case we do not have to rely on such methodological considerations. Instead, we have a beautiful example of a theory with unobservable “true time” – classical quantum theory. We have already mentioned that there is no self-adjoint operator for time measurement. As a consequence, no physical clock can provide a precise measure of time. There is always a small probability that a real clock will run backward with respect to it . Nonetheless, the time of quantum theory is not only an important part of quantum theory. A lot of people have tried hard to remove time from quantum theory, without success. The example of time in quantum theory is not only a powerful illustration of the failure of positivistic argumentation, but answers the question about the physical meaning of true time: the simple answer is “the same as in classical quantum theory”. The advantage of this answer is that we do not need any vague metaphysical considerations about the nature of “true time”. ### C.3 Relativity as an insight about true time On the other hand, positivistic argumentation against true time is not the only relativistic argumentation. It seems, a lot of relativistic scientists acknowledge very well that there is more behind the notion of time than simple clock measurement. Indeed, let’s for this argument accept the positivistic argumentation that there is nothing like “true time”, and the only physically meaningful notion of time is clock time. What, in this case, is the physical meaning of the following – rather typical – relativistic argumentation : > One often hears that what General Relativity did was to make time depend on gravity … Such a dependence of time on gravity would have been strange enough for the Newtonian view, but General Relativity is actually much more radical than that.<sup>12</sup><sup>12</sup>12We have to disagree with this claim that a dependence of apparent time on gravity would have been strange for the Newtonian view. Last not least, a classical clock based on Newtonian theory – the pendulum clock – obviously depends on gravity. … Rather the theory states that the phenomena we usually ascribe to gravity are actually caused by time’s flowing unequable from place to place…. Most people find it very difficult even to imagine how such a statement could be true. … That gravity could affect time, or rather could affect the rate at which clocks run, is acceptable, but that gravity is in any sense the same as time seems naively unimaginable. As we see, the point is not that apparent time – time measurement with clocks – is influenced by gravity. This is only the part which is already “acceptable” to people who have not really understood relativity. The point which is considered to be the “naively unimaginable” insight of relativity is the identification with “time” – obviously not “apparent time”, but “true time”. Thus, the non-existing ghost of true time revives here in its full beauty, as an important, fundamental insight of relativity. In my understanding we have here a conflict between two common relativistic argumentations. On one hand, the rejection of the notion of “true time” and the reduction of time to clock time measurement, on the other hand the metaphysical identification of the same “true time” with time measurement, presented as an “insight of relativity”. This contradiction shows that there is not much in support of this “insight” except methodological confusion: Or we accept that there is nothing behind “time” except clock measurement. Then no non-trivial insight into the nature of “true time” exists. The (nonetheless very important) insight of relativity is about the influence of gravity on clocks. Or we accept that time is more than clock measurement. Then there is no empirical base for the identification of clock time with this other notion of time. The identification is not only a purely metaphysical hypothesis without empirical support. Moreover, this metaphysical hypothesis is highly questionable, it seems to be based only on misunderstanding because the two notions are usually described with the same word – time. Simply two completely different notions of time which have been distinguished already by Newton have been mingled. ### C.4 Technical aspects of the problem of time The metaphysical choice to identify quantum true time with relativistic clock time leads to subsequent technical problems. In different approaches to quantum gravity the problem of time shows in different ways. It is not the purpose of this paper to consider them, we refer here to , , for further details. But an example may be informative. In the canonical approach which seems to be the closest to our approach, some “internal coordinates” have to be chosen. But this leads to a “multiple choice problem” : > Generically, there is no geometrically natural choice for the internal spacetime coordinates and, classically, all have an equal standing. However, this classical cornucopia becomes a real problem at the quantum level since there is no reason to suppose that the theories corresponding to different choices of time will agree. > > The crucial point is that two different choices of internal coordinates are related by a canonical transformation and, in this sense, are classically of equal validity. However, one of the central properties/problems of the quantization of any non-linear system is that, because of the well-known Van-Hove phenomenon , most classical canonical transformations cannot be represented by unitary operators while, at the same time, maintaining the irreducibility of the canonical commutation relations. This means that in quantizing a system it is always necessary to select some preferred sub-algebra of classical observables which is to be quantized. ## Appendix D Quantum gravity requires a common background In this section, we consider an interesting problem of the quantization of GR. This problem has been the starting point for the author. It strongly suggest that for successful quantization GR should be modified, especially that it is necessary to introduce a common background manifold. The main difference between GR and other theories including SR is that GR is a “one world theory”. It does not have a possibility to compare different gravitational fields. In other theories we have some common background. This allows to talk about values at “the same point” for different solutions. In GR, this is not possible. Another solution may be defined even on another manifold. This is not problematic because a possibility to compare different solutions is not necessary in the classical domain. As long as we consider classical gravitational fields, there is only one solution we can observe, there is only “one world”. Therefore, in the classical domain a “one world theory” does not cause problems. But what about quantum superposition? In a superpositional state we can consider different classical solutions in a single state. Is a “one world theory” sufficient to describe superpositions? Or is there more information hidden in a superpositional state? Especially, is there information about relative position between the states which are in superposition? This simple question may be the key for the understanding of quantum theory. There seem to be only two ways to answer this question: yes or no. The relativistic answer is a clear no. Our answer is a clear yes. A nice way to formalize the two possible answers has been proposed by Anandan : the notions of c-covariance and q-covariance. We consider superpositions of quasi-classical gravitational fields. Now, we have to look what happens if we apply diffeomorphisms. We have different fields, and this raises the question if we are allowed to use different diffeomorphisms for these different fields or not. This leads to two ways to generalize classical covariance: The first, weak generalization is c-covariance – it requires covariance only if we apply the same diffeomorphism. The second, strong generalization is q-covariance – it allows different diffeomorphisms for the different states in superposition. Now, the relativistic answer is unique – the only appropriate generalization is q-covariance. The reason is that c-covariance requires an essential modification of classical GR. Indeed, the Einstein equations of classical GR defines a pair of solutions only modulo q-covariance. The GR Lagrangian is also q-covariant. Thus, to define the evolution of c-covariant but not q-covariant objects we simply do not have appropriate equations in classical GR. Moreover, even the notion of “the same diffeomorphism” is not appropriately defined in GR. Different GR solutions live on different manifolds, and there is no well-defined notion of “the same diffeomorphism” on another manifold. This is the solution which has been proposed by Anandan too. We have chosen the other way. We think that the description of a superposition requires additional, non-q-covariant information. The purpose of this section is to justify this choice. Let’s start with a simple question – the superposition of a state with a shifted version of itself: $`|g_{ij}(x)+|g_{ij}(xx_0)`$. Do you really think this may be simply the original state, not a non-trivial superposition? I’m not. But, of course, this simple argument does not seem to be sufficient. Fortunately, we can present a much stronger argumentation. As we see, already a very simple scattering experiment on such superpositional states gives observables which depend on “relative position”. ### D.1 A non-relativistic quantum gravity observable In the consideration of quantum gravity, people usually consider two basic theories: classical general relativity and relativistic quantum field theory on a fixed background. But these are not really the theories which have to be unified – the gravitational field is classical in above theories. The interesting point of quantum gravity is, of course, the consideration of superpositions of gravitational fields. It is, last not least, superposition which makes quantum theory different from classical statistics. And this difference will be the point of our experiment: what we want to measure is the transition probability which decides if a superposition has been destroyed by measurement or not. The problem with semi-classical QFT is that it does not give a base for such considerations. On the other hand, it is not at all difficult to compute such transition probabilities in a reasonable approximation. For this purpose, we can use a well-known simple theory which allows to consider non-trivial superpositions of gravitational fields. This theory is simply non-relativistic quantum gravity – classical multi-particle Schrödinger theory with Newtonian interaction potential. There is nothing ill-defined with this theory. From theoretical point of view it works as well as multi-particle Schrödinger theory with Coulomb interaction potential. The fact that we have no data is not really problematic. Indeed, the theory unifies classical quantum principles with classical gravity in an ideal way. If somebody tends to doubt simply because there are no data, quantum gravity is a forbidden area for him. Thus, to assume the correctness of Schrödinger theory in the non-relativistic limit is not problematic. But, of course, it is a non-trivial decision: We assume that classical Schrödinger theory is the non-relativistic limit of quantum gravity. Now, based on this non-relativistic theory we can consider simple gravitational scattering. This gives some insight into features of superpositions of gravitational fields, even if the field itself is not quantized. We simply have to consider superpositional states of an otherwise neutral particle. We need the particle only as a source of gravity. Thus, let’s consider a situation like a double slit experiment, with some superpositional state $`|\psi ^1+|\psi ^2`$ of a source of gravity. For the interaction with a test particle $`|\phi `$ only the gravitational field of the source is important. Thus, it is possible to interpret the interaction also as an interaction of the test particle $`|\phi `$ with a superpositional state of the gravitational field $`|g^1+|g^2`$. Now, let’s make some simplifying assumptions. First, let the mass M of the source particle be much greater than the mass m of the test particle: $`Mm`$. In this case, the state of the heavy particle is much less influenced by the interaction than the state of the test particle. Let’s also assume that the state of the source particle is highly localized: $`\psi ^1(x)\delta (xx_1)`$, $`\psi ^2\delta (xx_2)`$. In this case, we can use single particle theory to compute the result of the interaction. Let’s denote with $`\phi ^1,\phi ^2`$ the solution of the Schrödinger equation for the source particle located in $`x_1`$ resp. $`x_2`$. Then, for the two-particle problem with the initial values $`|\psi ^1|\phi `$ resp. $`|\psi ^2|\phi `$ we obtain approximately a tensor product solution $`|\psi ^1|\phi ^1`$ resp. $`|\psi ^2|\phi ^2`$. For the superpositional state $`(|\psi ^1+|\psi ^2)|\phi `$ we obtain the solution by superposition: $$|\psi ^1||\phi ^1+|\psi ^2||\phi ^2.$$ But this is equivalent to $$(|\psi ^1+|\psi ^2)(|\phi ^1+|\phi ^2)+(|\psi ^1|\psi ^2)(|\phi ^1|\phi ^2)$$ Now, we are interested in the transition probability $`|\psi ^1+|\psi ^2|\psi ^1|\psi ^2`$. We obtain $$p_{trans}=\frac{1}{2}(1Re\phi ^1|\phi ^2)$$ To understand why we are very interested in this transition probability let’s consider the limiting cases: if $`\phi ^1|\phi ^2=1`$, gravitational interaction is not important, the position of the source particle does not influence the state of the test particle. Thus, the superpositional state remains unchanged, the interaction with the test particle was not a measurement of position of the source particle. We have a tensor product state. Therefore, we can ignore the test particle and obtain a pure one-particle state for the source. In the other limiting case, the resulting states are orthogonal, $`\phi ^1|\phi ^2=0`$, therefore, the transition probability is $`\frac{1}{2}`$. We do not have a product state. If we ignore the test particle, we do not obtain a pure source particle state. Instead, we obtain a classical mixed state: $$\frac{1}{2}(|\psi ^1\psi ^1|+|\psi ^2\psi ^2|)$$ Thus, the transition probability defines if the interaction was a measurement which has destroyed the superposition or not. If there is something which allows to distinguish a superposition from a classical mixed state, than this “something” gives us information about the transition probability. If not, then we cannot distinguish the superposition from a mixed state. ### D.2 The problem: generalization to relativistic quasi-classical gravity Therefore, this transition probability is very important. If it is not observable, what distinguishes the theory from a classical statistical theory? Is a theory which does not allow to distinguish a superposition from a classical mixed state worth to be named “quantum theory”? This seems questionable. Thus, the assumption that this transition probability remains observable in full, relativistic quantum gravity seems to be a very natural one. Now, to compute the transition probability, we need the scalar product $`\phi ^1|\phi ^2`$. For the derivation of this formula, we have used only very few fundamental principles. Therefore, it seems reasonable to assume that this formula may be generalized. We make the following hypothesis: The scalar product $`\phi ^1|\phi ^2`$ is well-defined in relativistic quantum gravity. Now, let’s consider how to generalize it into the relativistic domain. The basic states of the source particle $`\psi ^1(x)\delta (xx_1)`$ resp. $`\psi ^1(x)\delta (xx_1)`$ we generalize into relativistic gravitational fields $`g_{ij}^1(x)`$ resp. $`g_{ij}^2(x)`$. For $`\phi (x)`$ we have to solve now, instead of the classical Schrödinger equation, a similar wave equation on these background metrics. We ignore the related field-theoretical problems with particle creation and so on and assume that the field equations may be solved without problems. Thus, we obtain two solutions $`\phi ^1(x)`$ resp. $`\phi ^2(x)`$ for the two gravitational fields. Now let’s consider the computation of the scalar product $`\phi ^1|\phi ^2`$, using the following naive formula as a base: $$\phi ^1|\phi ^2=\overline{\phi }^1(x)\phi ^2(x)𝑑x^3$$ The point is that this integral simply cannot be defined from point of view of classical general relativity. Indeed, if $`g_{ij}^1(x)`$ is a solution of the Einstein equations, we may apply an arbitrary transformation of coordinates and obtain the same solution in other coordinates. Now, if we apply such a transformation to $`g_{ij}^1(x)`$, the same transformation has to be applied to $`\phi ^1(x)`$ too to obtain the same solution in the other coordinates. But nothing requires to apply the same transformation to $`g_{ij}^2(x)`$ and $`\phi ^2(x)`$. Now, if we apply a coordinate transformation to $`\phi ^1(x)`$, but not to $`\phi ^2(x)`$, the result of the integral changes in a completely arbitrary way. The same result may be formulated in another way: the functions $`\phi ^1(x)`$ resp. $`\phi ^2(x)`$ are defined on different manifolds – the manifolds defined by the spacetime metrics $`g_{ij}^1(x)`$ resp. $`g_{ij}^2(x)`$. And scalar products between functions on different manifolds are simply undefined. A third way to formulate this result is that this scalar product is only c-covariant but not q-covariant. Indeed, the integral does not change if we apply the same diffeomorphism to above configurations, but changes if we apply different diffeomorphisms to above configurations. Therefore, it cannot be observable in q-covariant quantum GR. A consequence of this situation is that GR is completely unable to make predictions for the scalar product, even if the scalar product is given for some initial values. Indeed, assume we have fixed some initial values $`g_{ij}^1(x)`$, $`\phi ^1(x)`$, $`g_{ij}^2(x)`$, $`\phi ^2(x)|_{x^0=0}`$ and an appropriate number of derivatives. Thus, for the initial values we have restricted the freedom of choice of diffeomorphisms. But this does not help: there are diffeomorphisms which are identical for the initial values, with all derivatives. And the Einstein equations define the solution only modulo arbitrary diffeomorphisms. What seems to be even more serious is that the problem appears already in the non-relativistic limit. In Schrödinger theory, the scalar product is well-defined. But the gravitational field may be as close as possible to the non-relativistic situation, the scalar product remains completely undefined in relativistic theory. It seems therefore highly problematic to obtain Schrödinger theory as the non-relativistic limit of a q-covariant theory of quantum gravity. ### D.3 The solution: a fixed space-time background Let’s look now how this problem is solved in GET. We have some well-defined Newtonian background which is common for all field configurations. The additional term in the Lagrangian does not only give some additional term in the Einstein equations, but breaks relativistic diffeomorphism invariance. We have four additional equations – the harmonic coordinate equations. They define the solution uniquely, not only modulo diffeomorphism. Of course, “uniquely” is also a relative notion, it means relative to the Newtonian background. We can use the covariant formulation of GET, with a covariant equation for the variables used to define the background. The conceptual difference is that this background is common for all solutions. The resulting quantum theory is not q-covariant, but only c-covariant. The relative position between a general field configuration and the common Newtonian background defines relative positions between different field configurations. This allows to define scalar products as well as the notion “the same diffeomorphism” which is necessary to define c-covariance. Are there other ways to define the evolution of these scalar products? No. By accepting the existence and observability of the scalar product we have de facto accepted a common background manifold with preferred coordinates for all semi-classical gravitational fields. Indeed, we need only a few number of simple and natural restrictions to obtain a common position measurement for all gravitational fields. Assume as before that we have two configurations of gravitational fields $`g_{ij}^1(x)`$, $`g_{ij}^2(x)`$ on manifolds $`M_1`$ resp. $`M_2`$ and some Hilbert space of appropriate wave functions $`H(M_1)`$ and $`H(M_2)`$. Now, we assume the element $`\phi ^1(x)H(M_1)`$ has well-defined “scalar products” with all elements of $`H(M_2)`$. But in this case it defines a linear functional on $`H(M_2)`$. A linear functional on $`H(M_2)`$ uniquely defines an element of $`H(M_2)`$. This construction defines a map $`H(M_1)H(M_2)`$. Now, this map seems to be the appropriate place to define additional natural requirements. First, we need transitivity. If there are three spaces, the map $`H(M_1)H(M_3)`$ should be the same as the composition $`H(M_1)H(M_2)H(M_3)`$. As a special case $`M_1=M_3`$ we obtain that the map $`H(M_2)H(M_1)`$ is the inverse of $`H(M_1)H(M_2)`$. Another property is that they are norm-preserving. This is required to have a consistent probability interpretation. Above restrictions may be justified in the same way we have justified the existence of the scalar product itself: these properties are fulfilled in Schrödinger theory. But, once we have a norm-preserving map $`H(M_1)H(M_2)`$, we can use it to transfer a measurement from $`M_2`$ to $`M_1`$. Especially, we can transfer the position measurement for the manifold $`M_2`$ to $`H(M_1)`$. Now, we can choose an arbitrary solution as a reference solution and transfer its position measurement to all other states as the common background. Thus, we obtain a common background manifold for all field configurations. We are de-facto back to the scheme we use in GET, with a fixed common space-time background. Thus, if we follow the relativistic paradigm and develop a q-covariant quantum theory of gravity, important observables of non-relativistic quantum gravity remain undefined. If we assume that they are well-defined, we have to reject the relativistic paradigm and to introduce a common background manifold into the theory. This consideration justifies the introduction of a fixed space-time background into GET. ### D.4 Comparison of Regge calculus and dynamical triangulation It is interesting to compare two well-known discrete approaches to quantum gravity – the Regge calculus and dynamical triangulations (DT) from point of view of scalar product. In the Regge calculus, we have a fixed grid and the geometry is described by the edge lengths of the grid. In contrast, in DT the edge lengths are fixed but the triangulation varies. Now, if we have some discrete functions defined for different geometries in the Regge calculus, we can define their scalar product without problem – we have the same grid as the base, therefore, for each point of one geometry we have a well-defined notion of “the same point” on the other geometry. Therefore, the Regge calculus is a nice example of a discrete c-covariant theory. Instead, in DT we do not have such a possibility. The scalar product between discrete functions on different triangulations is meaningless. Therefore, it is an example of a discrete q-covariant approach. Thus, according to the ideology presented here, DT should be the “correct” way to quantize geometry in a diffeomorphism-invariant way, but leads to problems with the classical limit, while it should be reverse for the Regge calculus. Indeed, the review shows exactly these properties of the two approaches: the Regge calculus with fixed grid contains steps of freedom which are unphysical from point of view of relativity, especially modes corresponding to general coordinate transformations, but it “possesses a weak field expansion in which contact can be made with continuum perturbation theory”. It is also noted that, at least for 2D, “the DT method affords a good prescription for regulating quantum gravity”. But “there is no weak coupling limit in which contact can be made with continuum perturbation theory. Indeed, the attractive feature of this formulation – that it is purely geometric, making no reference to coordinates and metric tensors also poses a problem; how do such classical quantities emerge from the model at large distance”. I hope, these remarks help to clarify my understanding of the role of the scalar product problem: it does not claim that q-covariant theories are impossible. Instead, DT provides an example of the regularized version of such a theory. It also does not claim that by accepting the existence of a scalar product we immediately end with ether theory: there is a certain difference between Regge QG and our ether approach, especially in our ether approach the grid is not fixed, but moves, and the position of the grid nodes are steps of freedom of the ether approach. The point is that a q-covariant approach has to be rejected because of the failure to define a scalar product between functions defined on different solutions, because such scalar products are necessary in the non-relativistic limit. ## Appendix E Realism as a methodological concept In § 15 we have not considered the EPR criterion of reality. There was no necessity for this, because we have considered it as part of common sense. To separate one of the common sense principles used in this proof and to name it “EPR criterion” is not necessary in our approach. Instead, it is part of the destruction strategy we have considered in section § 15.4. Unfortunately, this destruction strategy was already successful. Therefore, it seems necessary to consider the part of common sense which has been named “EPR criterion of reality” and questioned in more detail. ### E.1 Principles of different importance If we have a contradiction between theory and experiment, there are always different parts of the theory which may be blamed for the problem. But often it is not too difficult to find the critical part. Usually it is very helpful that different parts are not on the same level of fundamentality. We usually can distinguish more and less fundamental parts of the theory, and in case of conflict we usually blame the less fundamental parts to be the cause of the problem. Let’s consider, for example, the dark matter problem. We observe a difference between the Einstein equations and observation: $$G^{\mu \nu }8\pi GT_{obs}^{\mu \nu }$$ In this case, we do not reject GR because it’s main equation is falsified by observation. Instead, we simply define the energy-momentum tensor of “dark matter” as $$T_{dark}^{\mu \nu }=G^{\mu \nu }8\pi GT_{obs}^{\mu \nu }$$ and obtain that the Einstein equations are fulfilled. In this case, our existing theory of matter is considered to be less fundamental. In this case, this seems to be a reasonable choice. We argue here that the situation is different for the violation of Bell’s inequality. The other principles involved in the proof of Bell’s inequality, which we denote here as the principles of realism and which include the EPR criterion of reality, are more fundamental than the particular assumption about space-time symmetry known as relativity. Especially, we argue that these other principles are fundamental methodological principles, part of the methodological foundations of science. Thus, they are important as in physics, as in other sciences. Note that this argument is only additional support for our argumentation. There is already another argumentation which is completely sufficient for a unique decision in favor of realism: There is independent evidence against relativity – their problems with quantum gravity, especially the problem of time (appendix C). Moreover, there is a viable competitor of relativity – ether theory – which has been developed to solve these other problems. Instead, nobody has proposed a theory which, for independent reasons, rejects realism. Moreover, there is simply no independent evidence against realism: as we have seen, quantum theory is compatible with realism. ### E.2 A definition of reality and causal influence Let’s try to define classical realism in a way which allows a strong mathematical proof of Bell’s inequality based on this notion of realism. Realism in the common sense proposes the existence of an observer-external reality which exists independent of our observation. The results of observations may be results of complex interactions between reality and observer, nothing requires a possibility of direct observation. To define realism we need at least the following three entities: of course the observables, but also the decisions of the experimenters what to measure, and last not least the reality. But that’s all we need: ###### Axiom 1 (reality) Assume we have an experiment described by observables $`X`$ with the observable probability distribution $`\rho _X(X,x)dX`$. It depends on a set of control parameters $`x`$ which describe the experimental setup (the decisions of experimenters). A theory is realistic if it describes such probability distributions based on a notion of reality – a space $`\mathrm{\Lambda }`$ (reality) with probability distribution $`\rho _\lambda (\lambda )d\lambda `$ – and a realistic explanation – a function $`X(x,\lambda )`$ – so that for a test function f $$f(X)\rho _X(X,x)𝑑X=f(X(x,\lambda ))\rho (\lambda )𝑑\lambda $$ This formal definition is in quite good agreement with the common sense idea that reality $`\lambda `$ exists independent of our decisions $`x`$: the probability distribution $`\rho (\lambda )`$ indeed does not depend on $`x`$. But it already incorporates the insight that there is no pure observation, that our observations are only the result of complex interactions between observer and reality. An argumentation that classical realism is invalid because observations are only the result of such interactions is, therefore, invalid: this possibility is already part of classical realism. Note that already on this level, without any relation to space-time, we can define causal influences in a natural way: ###### Axiom 2 (causal influence) If in a realistic theory an observable X depends on a control parameter x in the realistic explanation $`X(x,\lambda )`$, then we have a causal influence of x on X. ### E.3 Bell’s inequality as a fundamental property These definitions are already sufficient for the proof of Bell’s inequality. In the case of Bell’s inequality, the control parameters are the questions: your question $`a`$ to Alice and your friends question b to Bob. The observables are their answers A and B. Using the definition of realism we obtain the existence of two functions: $$A=A(a,b,\lambda )=\pm 1;B=B(a,b,\lambda )=\pm 1$$ We also obtain the expectation value for the product AB as $$P(a,b)=\rho (\lambda )A(a,b,\lambda )B(a,b,\lambda )𝑑\lambda .$$ Now we have to consider causality. If there is no causal influence of the decision b on the result A, then we have $`A=A(a,\lambda )`$ and resp. $`B=B(b,\lambda )`$. Thus, we obtain $$P(a,b)=\rho (\lambda )A(a,\lambda )B(b,\lambda )𝑑\lambda $$ which is simply formula (2) of . After this, Bell’s inequality follows as derived in . As a consequence, we obtain the proof of Bell’s inequality on a level where even the existence of something like space-time has not been mentioned, and therefore in a space-time independent form: if it is violated, this proves the existence of causal influences $`(aB)`$ or $`(bA)`$. ### E.4 The methodological character of this definition A remarkable property of this definition is its unfalsifiable character. Whatever we observe, it is possible to describe it using a probability distribution $`\rho _X(X,x)dX`$. Whatever this probability distribution is, we can always construct a realistic theory which leads to this distribution – all we have to do is to use a (sufficiently artificial) functional space to describe the reality. Especially, reality may be described simply by the measure $`\rho _X(X,x)dX`$ itself.<sup>13</sup><sup>13</sup>13As the many worlds interpretation, as Bohmian mechanics may be considered as realistic theories obtained as variants of this “cheap” way – they simply accept the wave function as reality. The reasonable question is about the purpose of this definition if it is unfalsifiable. Now, there is a surprisingly simple answer: realism is simply a methodological rule. It enforces to describe certain parts of the theory as really existing. As well, the subsequent definition of a causal influence is also unfalsifiable. The purpose of this definition is, as well, to enforce to name some relations causal influence. In other words, this definition of realism and causality enforces ontological clarity. A realistic theory is a theory where we are forced to name some things real and some influences causal, even if this violates our metaphysical prejudices or principles of the theory we prefer. Especially this definition of realism and causality, a definition already enforces that any violation of Bell’s theorem should be explained by causal influences – or $`aB`$ or $`bA`$. ### E.5 Causality requires a preferred frame It does not follow from the definition that these causal influences happen in a preferred frame. To prove the existence of a preferred frame we need a little bit more. First, the connection between causality and space-time. Until now, even the existence of something like a space-time has not been mentioned. Only now we have to define causality on space-time as a relation $`xy`$ between space-time events x, y: $`xy`$ if there exists some $`aA`$ so that the decision a is localized at x and the observation A localized at y. Moreover, we need the most important property of causality: the causal order along a world line and the absence of causal loops. ###### Axiom 3 (space-time causality) Causality defines a partial order $`xy`$ on space-time with the property that on time-like trajectories $`\gamma (t)`$ we have $`\gamma (t_0)\gamma (t_1)`$ if $`t_0<t_1`$. Now, we can simply prove the existence of a preferred foliation: <sup>14</sup><sup>14</sup>14That the existence of a preferred foliation follow is nothing new. Valentini suggests “that a preferred foliation of spacetime could arise from the existence of nonlocal hidden-variables” . Bell himself concludes : “the cheapest resolution is something like going back to relativity as it was before Einstein, when people like Lorentz and Poincare thought that there was an aether — a preferred frame of reference — but that our measuring instruments were distorted by motion in such a way that we could no detect motion through the aether.” ###### Theorem 8 (existence of a preferred foliation) If for all pairs of events Bell’s inequality may be violated, then there exists a preferred foliation. It is defined by the property that if $`xy`$ than $`T(x)<T(y)`$ for the function T(x) which defines this foliation. Proof: Let’s define the foliation as a time-like function $`T(x^i,t)`$. For this purpose, we set $`T(0,t_0)=t_0`$ and define the points contemporary to $`A=(0,0,0,t_0)`$ on the line $`B(t)=(x^1,x^2,x^3,t)`$. We use the Dirichlet algorithm on the line $`B(t)`$. We start with a large enough interval $`(a_0,b_0)`$ so that there exists causal influences $`B(a_0)AB(b_0)`$. Assume at step n we have found an interval $`B(a_n)AB(b_n)`$ with $`|b_na_n|<2^n|b_0a_0|`$. Now, we consider the element $`B(h=(b_n+a_n)/2)`$. Then we observe a violation of Bell’s inequality between A and $`B(h)`$. It follows from Bell’s theorem that there should be $`AB(h)`$ or $`B(h)A`$. In the first case, we set $`a_{n+1}=a_n,b_{n+1}=h`$, else $`a_{n+1}=h,b_{n+1}=b_n`$. In above cases we have found an interval with $`B(a_{n+1})AB(b_{n+1})`$ with $`|b_{n+1}a_{n+1}|<2^{n1}|b_0a_0|`$. Therefore, we have a limit $`l=lima_n=limb_n`$. This limit defines a function $`T(x^1,x^2,x^3,t_0)=l`$. To prove that the function $`T(x)`$ is correctly defined, Lipschitz continuous, and that the definition of the foliation does not depend on the choice of coordinates is straightforward. What we need is that in every environment of $`B(l)`$ we have points $`a_n`$ with $`B(a_n)A`$ as well as points $`b_n`$ with $`AB(b_n)`$, the non-existence of causal loops, and the existence of causal ordering $`B(a)B(b)a<b`$ on time-like trajectories $`B(t)`$. ### E.6 Relation between our definition and the EPR criterion Let’s consider now the difference between this definition of realism and the EPR criterion of reality : > If, without in any way disturbing a system, we can predict with certainty … the value of a physical quantity, than there exists an element of physical reality corresponding to this physical quantity. Now, this is a natural consequence of our definition of realism: We have no “disturbance”, thus, no dependence of A on b: $`A=A(a,\lambda )`$. If it is a prediction, because we have no influence backward in time, we have no dependence of B on a: $`B=B(b,\lambda )`$. It is a prediction with certainty, thus, these functions are identical as functions: $`A(.,\lambda )=B(.,\lambda )`$. Last not least, $`\lambda `$ is the “element of reality” and the function $`A(.,\lambda )`$ describes the correspondence to the physical quantity A. The advantages of our new definition are, in our opinion, the following: we have a general definition, while the EPR mentions a special situation – a correlation which allows to predict something with certainty, and we do not depend on the notion of space-time, while the EPR criterion includes an implicit reference to time (“predict”). Moreover, the formal character of the definition allows to show its methodological character: it does not restrict physical theories, but restricts our way to talk about them. ### E.7 Methodological principles as the most fundamental part of science Once we defend realism as a fundamental methodological principle it seems useful to look how other fundamental principles of science may be defended. There is another such fundamental methodological principle – classical logic, especially the law that there should be no contradictions, nor in the theory, nor between theory and observation. As our definition of realism and causality these principles are unfalsifiable themself – simply because the principle of falsification itself relies on classical logic.<sup>15</sup><sup>15</sup>15If they are false, then the method of falsification is false too, therefore, cannot be used. Even if the use of rational arguments, especially the “therefore” in the last sentence, is also unjustified, this argument seems to show that an experimental falsification of classical logic is impossible. Therefore, other arguments have to be used to defend them. In this context, it is interesting how Popper defends classical logic against “dialectical logic” (, p.316): > Dialecticians say that contradictions are fruitful, or fertile, or productive of progress, and we have admitted that this is, in a sense, true. It is true, however, only so long as we are determined not to put up with contradictions, and to change any theory which involves contradictions; in other words never to accept a contradiction: it is solely due to this determination of ours that criticism, i.e. the pointing out of contradictions, induces us to change our theories, and thereby to progress. It cannot be emphasized too strongly that if we change this attitude, and decide to put up with contradictions, then contradictions must at once lose any kind of fertility. Thus, the point of the argumentation is not to prove that there can be no contradictions. The basic idea is that we have to consider not the hypothesis itself, but their influence on the future development of science. That means, Popper defends classical logic as a methodological rule of science. The main advantage of this argumentation is that we do not have to rely on “common sense” – a notion which has a bad name in current science and is usually compared with flat Earth theory. Classical logic is not a particular common sense theory like flat Earth theory which may be false, but defines the scientific method, therefore, if we reject classical logic, we simply reject the scientific method. Our notion of realism is fertile in the same sense as classical logic. It is the rule to search for realistic, causal explanations for observable correlations. A non-trivial, unexplained correlation plays the same role as the contradiction in logic: it defines a scientific problem. We have to include a realistic explanation into our theory. Nobody forces us to search for explanations, it is only our own methodological decision not to accept unexplained correlations, and to accept only a realistic explanation. If we give up the search for realistic explanations, we loose an important way to reach scientific progress. ### E.8 The methodological role of Lorentz symmetry The great importance of Lorentz symmetry in modern physics is often presented as if it is a decisive argument against a preferred frame. But this suggests that Lorentz symmetry would have been less important in the Lorentz ether. Is there any evidence for this claim? I have never seen any justification for this assumption. It is simply claimed, without justification, that people would have been less eager to search for relativistic symmetry. The reverse may be closer to truth. Instead, with the Lorentz ether as the leading ideology, people would have tried to detect hidden Lorentz symmetry in usual condensed matter theory. Moreover, without doubt any part of the hidden variables which can be made Lorentz-covariant would have been made Lorentz-covariant. An example are the equations of GET presented here. The new equation for the preferred coordinates is a nice, well-known relativistic equation – the harmonic equation. Moreover, the thesis is in obvious contradiction to the history of the Lorentz ether. In the context of the Lorentz ether, by Poincare, the program to make all physical theories Lorentz-invariant has been proposed in general and realized for kinematics. The decision to reject the existence of a preferred frame made by Einstein was in no way necessary for the development of this program. The point is that to require a particular symmetry is not a general methodological rule of scientific research. At best there is the related methodological rule to search for symmetries in general. But even this rule seems much less fundamental than classical logic and realism: last not least, we search for symmetries in reality. Symmetries are a powerful tool to study realistic theories, to detect contradictions in such theories or between theory and experiment – but only a tool, in no way a fundamental principle. ### E.9 Discussion As presented here, the preferred frame is the unavoidable consequence of the violation of Bell’s inequality. Relativity is falsified by Aspect’s experiment, and its current status should be rejected as an immunization. This is so obvious that it becomes problematic to explain the unreasonable decision of mainstream science to reject realism. But there are several factors which may be blamed here: * The absence of a reasonable theory of gravity with preferred frame. This problem is solved now by GET. * The widely accepted belief, based on von Neumann’s theorem, that hidden variable theories are impossible. This was justified at the time the EPR criterion was proposed, but many seem to believe it even today. * The ignorance of Bohmian mechanics because it requires a preferred frame. * The extreme positivism and subjectivism during the foundational period of quantum theory. * The general ignorance of fundamental problems of quantum theory today. But the most important explanation seems to be Kuhn’s theory of paradigm shifts . According to Kuhn, paradigms are never falsified by experiments. A paradigm may be rejected only if a new paradigm appears. Until now, no alternative paradigm has been proposed, therefore, to preserve the relativistic paradigm was justified – in full agreement with Kuhn’s paradigm shifts. We propose here a new paradigm – a return to classical Newtonian space-time and ether theory. With this paradigm as a competitor of the relativistic paradigm it is no longer necessary to reject realism or causality. ## Appendix F Bohmian mechanics An essential property of non-relativistic Schrödinger theory is the existence of a simple deterministic interpretation – Bohmian mechanics (BM). We refer to this theory in our proof that EPR realism is not in conflict with non-relativistic quantum theory. Unfortunately, BM is widely ignored. The main reason for this ignorance seems to be that it requires a preferred frame – thus, a feature which makes it particularly attractive in the context of GET. Therefore, it seems reasonable to consider the basic features of BM here. ### F.1 Simplicity of Bohmian mechanics BM may be considered as a straightforward way to complete quantum mechanics. In BM, we have two entities: the “guiding wave” $`\mathrm{\Psi }(q)`$ defined on the configuration space which fulfills the classical Schrödinger equation $$i_t\mathrm{\Psi }=H\mathrm{\Psi }$$ and the configuration $`Q(t)`$ which fulfills the so-called “guiding equation”. This guiding equation may be obtained in a straightforward way from quantum mechanics. The basic observation is the following: quantum mechanics provides us with a probability current $`j^i(q)`$ as well as with a probability density $`\rho (q)=\mathrm{\Psi }^{}(q)\mathrm{\Psi }(q)`$. In classical mechanics they are related by $`j^i(q)=\rho (q)v^i(q)`$. Now it requires no great imagination to write the guiding equation $$\frac{dQ}{dt}=v^i=\frac{j^i}{\rho }$$ This defines the evolution of the state. Now, if in initially the state is in the so-called “quantum equilibrium” $`\rho (q)`$, then it remains in this state. This follows from the continuity equation $$_t\rho (q,t)+_ij^i(q,t)=0$$ That’s already all what is necessary. There is no need for further axioms. Therefore, all we need for the definition of Bohmian mechanics is the quantum probability current. For example, in non-relativistic multi-particle theory $$H=\underset{k=1}{\overset{N}{}}\frac{\mathrm{}^2}{m_k^2}_k^2+V(q_1,\mathrm{},q_N)$$ this probability current is given by $$j_k=\frac{\mathrm{}}{m_k}\mathrm{}(\psi ^{}_k\psi )$$ Therefore, we obtain the guiding equation $$\frac{dQ_k}{dt}=\frac{\mathrm{}}{m_k}\mathrm{}\frac{_k\psi }{\psi }$$ ### F.2 Clarity of the interpretation The first thing we have to note here is the ontological clarity. To quote Bell (, p.191): > Is it not clear from the smallness of the scintillation on the screen that we have to do with a particle? And is it not clear, from the diffraction and interference patterns, that the motion of the particle is directed by a wave? … This idea seems to me so natural and simple, to resolve the wave-particle dilemma in such a clear and ordinary way, that it is a great mystery to me that it was so generally ignored. … This solution of the wave-particle confusion not the main point: in BM there is also nothing strange with Schrödinger’s cat. The wave function of the cat remains in its superpositional state, but the actual cat is in a well-defined state. “There is no need in this picture to divide the world into ‘quantum’ and ‘classical’ parts. For the necessary ‘classical terms’ are available already for individual particles (their actual positions) and so also for macroscopic assemblies of particles.” (, p.192) Another interesting question is worth to be mentioned: why are the states we observe in quantum equilibrium? This question has an interesting answer: decoherence. “One of the best descriptions of decoherence, though not he word itself, can be found in Bohm’s 1952 ‘hidden variables’ paper . We wish to emphasize, however, that while decoherence plays a crucial role in the very formulation of the various interpretations of quantum theory loosely called decoherence theories, its role in Bohmian mechanics is of quite different character: For Bohmian mechanics decoherence is purely phenomenological – it plays no role whatsoever in the formulation (or interpretation) of the theory itself” . The most important property of BM is it’s compatibility with classical principles: the EPR criterion of reality, classical causality, determinism. Let’s quote again Bell (, p.163): > It is easy to find good reasons for disliking the de Broglie-Bohm picture. Neither de Broglie nor Bohm liked it very much; for both of them it was only a point of departure. Einstein also did not like it very much. He found it ‘too cheap’, although, as Born remarked, ‘it was quite in line with his own ideas’. But like it or lump it, it is perfectly conclusive as a counter example to the idea that vagueness, subjectivity, or indeterminism, are forced on us by the experimental facts covered by non-relativistic quantum mechanics. ### F.3 Relativistic generalization Let’s consider now the main point why many researchers dislike BM – its relativistic generalization. The same basic scheme works as well in relativistic theory and field theory. For example, for multiple Dirac particles Bohm has proposed the following guiding equation: $$𝐯_k=\frac{\psi ^+\alpha _k\psi }{\psi ^+\psi }$$ For the general case of quantum field theory, we have to accept the lectures of quantum field theory what is the appropriate notion of the wave function: “Certainly the Maxwell field is not the wave function of the photon, and for reasons that Dirac himself pointed out, the Klein-Gordon fields we use for pions and Higgs bosons could not be the wave functions of the bosons. In its mature form, the idea of quantum field theory is that quantum fields are the basis ingredients of the universe, and particles are just bundles of energy and momentum of the fields. In a relativistic theory the wave function is a functional of these fields, not a function of particle coordinates” . Thus, it does not make sense to search for a guiding equations for particles in the general case, and we have to consider Bohmian field theory where we obtain a guiding equation for generalized coordinates – the field configuration. Thus, the generalization itself is not problematic. It is an essential property of this generalization – that it has an explicit preferred frame on the fundamental level. The predictions are nonetheless Lorentz-invariant. For example, $`\psi ^+\psi `$ is an equivariant ensemble density in the chosen reference frame. It reproduces the quantum predictions in this frame. These predictions don’t contain a trace of the preferred frame. Lorentz invariance holds on the observational, but not on the fundamental level. The 4-tuple $`(\psi ^+\psi ,\psi ^+\alpha _k\psi )`$ is not a 4-vector for $`N>1`$. This is not an accident. “There does not in general exist a probability measure P on N-paths for which the distribution of crossing $`\rho ^\mathrm{\Sigma }`$ agrees with the quantum mechanical distribution on all space-like hyper-planes $`\mathrm{\Sigma }`$. This assertion is a more or less immediate consequence of Bell’s inequality: by means of a suitable placement of appropriate Stern-Gerlach magnets the inconsistent joint spin correlations can be transformed to (the same) inconsistent spatial correlations for particles at different times . Thus, we have the probability measure $`\rho =|\psi |^2`$ only in one frame. But this measure in just one frame is sufficient to derive the quantum mechanical predictions for observations at different times. ### F.4 Discussion The fact that Lorentz invariance does not hold on the fundamental level is often considered as a decisive argument against BM. But from point of view of ether theory this becomes a virtue rather than a vice: every argument in favour of BM becomes an argument in favour of the preferred frame we use in ether theory. To use the argument “there is no fundamental Lorentz-invariance” against BM in this context would be simply circular reasoning – a main concept of ether theory is as well that there is no fundamental Lorentz-invariance. Thus, BM gives additional support for one of the main ingredients of GET – the preferred absolute time. One the other hand, GET gives support to BM – it shows a way to generalize BM to gravity. We do not have to try to find Lorentz-invariant versions of BM, as tried, for example, in . Instead, we can apply BM as it is, with a preferred frame, in GET or, even better, in an atomic ether theory.
warning/0001/math0001111.html
ar5iv
text
# Redundant Picard–Fuchs system for Abelian integrals ## 1. Tangential Hilbert Sixteenth Problem, complete Abelian integrals and Picard–Fuchs equations The main result of this paper is an explicit derivation of the Picard–Fuchs system of linear ordinary differential equations for integrals of polynomial 1-forms over level curves of a polynomial in two variables, regular at infinity. The explicit character of the construction makes it possible to derive upper bounds for the coefficients of this system. In turn, application of the bounded meandering principle to the system of differential equations with bounded coefficients allows to produce upper bounds for the number of complex isolated zeros of these integrals on a positive distance from the ramification locus. ### 1.1. Abelian integrals and tangential Hilbert 16th problem If $`H(x,y)`$ is a polynomial in two real variables, called the Hamiltonian, and $`\omega =P(x,y)dx+Q(x,y)dy`$ a real polynomial 1-form, then the problem on limit cycles appearing in the perturbation of the Hamiltonian equation, $$dH+\epsilon \omega =0,\epsilon (,0)$$ (1.1) after linearization in $`\epsilon `$ (whence the adjective “tangential”) reduces to the study of complete Abelian integral $$I(t)=I(t;H,\omega )=_{H=t}\omega ,$$ (1.2) where the integration is carried over a continuous family of (real) ovals lying on the level curves $`\{H=t\}`$. ###### Problem 1 (Tangential Hilbert 16<sup>th</sup> problem). Place an upper bound for the number of real zeros of the Abelian integral $`I(t;H,\omega )`$ on the maximal natural domain of definition of this integral, in terms of $`\mathrm{deg}H`$ and $`\mathrm{deg}\omega =\mathrm{max}(\mathrm{deg}P,\mathrm{deg}Q)+1`$. A more natural version appears after complexification. For an arbitrary complex polynomial $`H(x,y)`$ having only isolated critical points, and an arbitrary complex polynomial 1-form $`\omega `$, the integral (1.2) can be extended as a multivalued analytic function ramified over a finite set of points (typically consisting of critical values of $`H`$). The problem is to place an upper bound for the number of isolated complex roots of any branch of this function, in terms of $`\mathrm{deg}H`$ and $`\mathrm{deg}\omega `$. ### 1.2. Abelian integrals and differential equations Despite its apparently algebraic character, the tangential Hilbert problem still resists all attempts to approach it using methods of algebraic geometry. Almost all progress towards its solution so far was based on using methods of analytic theory of differential equations. In particular, the (existential) general finiteness theorem by Khovanskiĭ–Varchenko claims that for any finite combination of $`d=\mathrm{deg}\omega `$ and $`n=\mathrm{deg}H`$ the number of isolated zeros is indeed uniformly bounded over all forms and all Hamiltonians of the respective degree. One of the key ingredients of the proof is the so called Pfaffian elimination, an analog of the intersection theory for varieties defined by Pfaffian differential equations . Another important achievement, an explicit upper bound for the number of zeros in the elliptic case when $`H(x,y)=y^2+p(x)`$, $`\mathrm{deg}p=3`$ and forms of arbitrary degree, due to G. Petrov , uses the fact that the elliptic integrals $`I_k(t)=x^{k1}y𝑑x`$, $`k=1,2`$, in this case satisfy an explicit system of linear first order system of differential equations with rational coefficients. This method was later generalized for other classes of Hamiltonians whose level curves are elliptic (i.e., of genus $`1`$), see and references therein. The ultimate achievement in this direction is a theorem by Petrov and Khovanskii, placing an asymptotically linear in $`\mathrm{deg}\omega `$ upper bound for the number of zeros of arbitrary Abelian integrals, with the constants being uniform over all Hamiltonians of degree $`n`$ (unpublished). However, one of these constants is purely existential: its dependence on $`n`$ is totally unknown. It is important to remark that all the approaches mentioned above, require a very basic and easily obtainable information concerning the differential equations (their mere existence, types of singularities, polynomial or rational form of coefficients, in some cases their degree). ### 1.3. Meandering of integral trajectories A different approach suggested in consists in an attempt to apply a very general principle, according to which integral trajectories of a polynomial vector field (in $`^n`$ or $`^n`$) have a controllable meandering (sinuosity), . More precisely, if a curve of known size is a part of an integral trajectory of a polynomial vector field whose degree and the magnitude of the coefficients are explicitly bounded from above, then the number of isolated intersections between this curve and any affine hyperplane in the ambient space can be explicitly majorized in terms of these data. The bound appears to be very excessive: it is polynomial in the size of the curve and the magnitude of the coefficients, but the exponent as the function of the degree and the dimension of the ambient space, grows as a tower (iterated exponent) of height 4. In order to apply this principle to the tangential Hilbert problem, we consider the curve parameterized by the monomial integrals, $$t(I_1(t),\mathrm{},I_N(t)),I_i(t)=_{H=t}\omega _i,$$ where $`\omega _i`$, $`i=1,\mathrm{},N`$ are all monomial forms of degree $`d`$. Isolated zeros of the Abelian integral of an arbitrary polynomial 1-form $`\omega =_ic_i\omega _i`$ correspond to isolated intersections of the above curve with the hyperplane $`c_iI_i=0`$. If this monomial curve is integral for a system of polynomial differential equations with explicitly bounded coefficients, then the bounded meandering principle would yield a (partial) answer for the tangential Hilbert 16th problem. The system of polynomial (in fact, linear) differential equations can be written explicitly for the case of hyperelliptic integrals corresponding to the Hamiltonian $`H(x,y)=y^2+p(x)`$ with an arbitrary univariate potential $`p(x)[x]`$, see §2 below and references therein. Application of the bounded meandering principle allowed us to prove in that the number of zeros of hyperelliptic integrals is majorized by a certain tower function depending only on the degrees of $`n=\mathrm{deg}H=\mathrm{deg}p`$ and $`d=\mathrm{deg}\omega `$. (Actually, it was done under an additional assumption that all critical values of $`p`$ are real, but we believe that this restriction is technical and can be removed). ### 1.4. Picard–Fuchs equations and systems of equations In order to generalize the construction from for the case of arbitrary (not necessarily hyperelliptic) Hamiltonians it is necessary, among other things, to write a system of polynomial differential equations for Abelian integrals and estimate explicitly the magnitude of its coefficients. The mere existence of such a system is well known since times of Riemann if not Gauss. In today’s language, the monodromy group of any form depends only on the Hamiltonian. Denote by $`\mu `$ the rank of the first homology group of a typical affine level curve $`\{H=t\}^2`$. Then for any collection of 1-forms $`\omega _1,\mathrm{},\omega _\mu `$ the period matrix $`X(t)`$ can be formed, whose entries are integrals of $`\omega _i`$ over the cycles $`\delta _1(t),\mathrm{},\delta _\mu (t)`$ generating the homology. If the determinant of this matrix if not identically zero, then $`X(t)`$ satisfies a linear ordinary differential equation of the form $$\dot{X}(t)=A(t)X(t),A()\mathrm{Mat}_{\mu \times \mu }((t)),$$ (1.3) with a rational matrix function $`A(t)`$. This system of equations is known under several names, from Gauss–Manin connection \[20, especially p. 18\] to Picard–Fuchs system (of linear ordinary differential equations with rational coefficients, in full). We shall systematically use the last name. The rank of the first homology can be easily computed: for a generic Hamiltonian of degree $`n+1`$ it is equal to $`n^2`$. The degree $`\mathrm{deg}A(t)`$ can be relatively easily determined if the degrees of the forms $`\omega _i`$ are known. However, the choice of the forms $`\omega _i`$ may also be a difficult problem for some Hamiltonians. The matrix $`A(t)`$ apriori may have poles not only in the ramification points of the Abelian integrals, which leads to additional difficulties. But worst of all, this topological approach gives absolutely no control over the magnitude of the (matrix) coefficients of the rational (matrix) function $`A(t)`$. ### 1.5. Regularity at infinity and Gavrilov theorems Part of these problems problems can be resolved. In particular, if the Hamiltonian is sufficiently regular at infinity, then all questions concerning the degrees, can be answered. ###### Definition 1. A polynomial $`H(x,y)[x,y]`$ of degree $`n+1`$ is said to be regular at infinity, if one of the three equivalent conditions holds: 1. its principal homogeneous part $`\widehat{H}`$, a homogeneous polynomial of degree $`n+1`$, is a product of $`n+1`$ pairwise different linear forms; 2. $`\widehat{H}`$ has an isolated critical point (necessarily of multiplicity $`\mu =n^2`$) at the origin $`(x,y)=(0,0)`$; 3. the level curve $`\{\widehat{H}=1\}^2`$ is nonsingular. This condition means that after the natural projective compactification of the $`(x,y)`$-plane $`^2`$, all “interesting” things still happen only in the finite part of the compactified plane. In particular, for a polynomial regular at infinity: 1. all level curves $`\{H=t\}`$ intersect the infinite line $`P_{\mathrm{}}^1P^2`$ transversally, 2. all critical points $`\{(x,y):dH(x,y)=0\}`$ are isolated and their number is exactly $`\mu =n^2`$ if counted with multiplicities, 3. the rank of the first homology of any regular affine level curve $`\{H=t\}`$ is $`\mu =n^2`$, 4. the map $`H:^2^1`$ is a topological bundle over the set of the regular values of $`H`$, hence the Abelian integrals can be ramified only over the critical values of $`H`$. In L. Gavrilov proved that for polynomials regular at infinity, the space of Abelian integrals is finitely generated as a $`[t]`$-module by $`\mu `$ basic integrals that can be chosen as integrals of any $`\nu `$ forms $`\omega _i`$ of degree $`2n`$ whose differentials form the basis of the quotient space $`\mathrm{\Lambda }^2/d\widehat{H}\mathrm{\Lambda }^1`$, where $`\mathrm{\Lambda }^k`$ is the space of polynomial $`k`$-forms on $`^2`$. As a corollary, one can prove that the collection of these basic integrals satisfies a system of equations (1.3) of size $`\mu \times \mu `$ with $`\mu =n^2`$, and place an upper bound for the degree of the corresponding matrix function $`A(t)`$. This system is minimal (irredundant): generically (for Morse Hamiltonians regular at infinity), all branches of full analytic continuation of an Abelian integral span exactly $`\mu `$-dimensional linear space. From this theorem one can also derive further information concerning the Picard–Fuchs system. Namely, one can prove that if in addition to being regular at infinity, $`H`$ is a Morse function on $`^2`$, then the matrix $`A(t)`$ of the Picard–Fuchs system (1.3) has only simple poles (Fuchsian singularities) at the critical values of the Hamiltonian and only at them (the point $`t=\mathrm{}`$ is a regular though in general non-Fuchsian singularity). However, these results do not yet allow an explicit majoration of the coefficients (e.g., the residue matrices) of the matrix function $`A(t)`$ in (1.3). ### 1.6. Redundant Picard–Fuchs system: the first main result We suggest in this paper a procedure of explicit derivation of the Picard–Fuchs system of equations, based on the division by the gradient ideal $`H_x,H_y[x,y]`$ in the polynomial ring. It turns out that if instead of choosing $`\mu =n^2`$ forms of degree $`2n`$ constituting a basis modulo the gradient ideal, one takes all $`\nu =n(2n1)`$ cohomologically independent monomial forms of degree $`2n`$, then the resulting Picard–Fuchs system can be written in the form $$(tEA)\dot{X}(t)=BX(t),A,B\mathrm{Mat}_{\nu \times \nu }(),$$ (1.4) where $`E`$ is the identity matrix, and $`X(t)`$ is the rectangular period $`\nu \times \mu `$-matrix. The procedure of deriving the system (1.4), being completely elementary, can be easily analyzed and upper bounds for the matrix norms $`A`$ and $`B`$ derived. These bounds depend on the magnitude of the all non-principal terms $`H\widehat{H}`$ of the Hamiltonian, relative to the principal part $`\widehat{H}`$. More precisely, we introduce a normalizing condition (quasimonicity) on the homogeneous part: this condition plays the same role as the assumption that the leading term has coefficient $`1`$ for univariate polynomials. The quasimonicity condition can be always achieved by an affine change of variables, provided that $`H`$ is regular at infinity, hence it is not restrictive. Theorem 2, our first main result, allows to place an upper bound for the norms $`A+B`$ in terms of the norm (sum of absolute values of all coefficients) of the non-leading part $`H\widehat{H}`$, assuming that $`H`$ is quasimonic. ### 1.7. Corollaries: theorems on zeros The above information on coefficients of Picard–Fuchs system already suffices to apply the bounded meandering principle and obtain an explicit upper bound for the number of zeros of complete Abelian integrals away from the critical locus of the Hamiltonian (Theorem 3), which seems to be the first known explicit result of that kind. In addition to this bound valid for some zeros and almost all Hamiltonians, one can apply results (or rather methods) from . If in addition to the quasimonicity and bounded lower terms, all critical values $`t_1,\mathrm{},t_\mu `$ of the Hamiltonian $`H`$ are far away from each other (i.e., a lower bound for $`|t_it_j|`$ is known for $`ij`$), then one can majorize the number of zeros on any branch of the Abelian integral by a function depending only on $`n,d`$ and the minimal distance between critical values. The accurate formulation is given in Theorem 4. ### 1.8. Equivariant formulation However, the description given by Theorem 2, is not completely sufficient for further advance towards solution of the tangential Hilbert problem by studying zeros of Abelian integrals near the critical locus when the latter (or some part of it) shrinks to one point of high multiplicity. One reason is that in order to run an inductive scheme similar to that constructed in , one has to make sure that the Hamiltonian $`H:^2^1`$ can be rescaled (using affine transformations in the preimage $`^2`$ and the image $`^1`$) so that simultaneously: 1. the critical values of $`H`$ do not tend to each other (e.g., their diameter is bounded from below by $`1`$), and 2. the “non-homogeneous part” $`H\widehat{H}`$ is bounded by a constant explicitly depending on $`n`$ (each of the two conditions can be obviously satisfied separately). Another, intrinsic reason is the equivariance (or, rather precisely, non-invariance of neither Theorem 2 nor Theorems 3 and 4) by the above affine group action. In order to be geometrically sound, all assertions should be related to a certain privileged affine chart on the $`t`$-plane. Since our future goal is to study a neighborhood of the critical locus, it is natural to choose the privileged chart so that the critical locus will not shrink into one point. More detailed explanations and motivations are given in §4 below, where we formulate several problems all in the following sense: for a polynomial whose principal homogeneous part is normalized (in a certain sense) and whose critical values are explicitly bounded, it is required to place an upper bound for the “non-homogeneous” part, eventually after a suitable translation (which does not affect the principal part, naturally). ### 1.9. Geometry of critical values of polynomials The reason why several problems of the above type were formulated instead of just one, is very simple: we do not know a complete solution, so partial, existential or limit cases were considered as intermediate steps towards the ultimate goal. In §5 we prove that: * if a monic complex polynomial $`p(x)=x^{n+1}+\mathrm{}[x]`$ has all critical values in the unit disk, then its roots form a point set of diameter $`<11`$ (Theorem 6) and hence by a suitable translation the norm of the non-principal part can be made $`12^{n+1}`$ (this gives a complete solution in the univariate and hyperelliptic cases); * all critical values of a Hamiltonian regular at infinity, cannot simultaneously coincide unless the Hamiltonian is essentially homogeneous (Theorem 5); * for any normalized principal part $`\widehat{H}`$ there exists an upper bound for $`H\widehat{H}`$ (eventually after a suitable translation), provided that the critical values of $`H`$ are all in the unit disk (Corollary to Theorem 5). All these are positive results towards solution of the problem on critical values. It still remains to compute the upper bound from the last assertion explicitly: the proof below does not provide sufficient information for that. However, it can be shown already in simple examples that this bound cannot be uniform over all homogeneous parts. As some of the linear factors from $`\widehat{H}`$ approach too closely to each other, an explosion occurs and the non-principal part may be arbitrarily large without affecting the “moderate” critical values. The phenomenon can be seen as “almost occurrence” of atypical values, ramification points for Abelian integrals that are not critical values of $`H`$: such points are known to appear if the principal part $`\widehat{H}`$ has a non-isolated singularity. ### Acknowledgements We are grateful to J.-P. Françoise, L. Gavrilov, Yu. Ilyashenko, A. Khovanskii, P. Milman, R. Moussu, R. Roussarie and Y. Yomdin for numerous discussions and many useful remarks. Bernard Teissier suggested an idea that finally developed into the proof of Lemma 3 below. Lucy Moser provided us with a counterexample (§5.6 below). We are grateful to all our colleagues from Laboratoire de Topologie, Université de Bourgogne (Dijon) and Departamento de Algebra, Geometría y Topología, Universidad de Valladolid, where a large part of this work has been done. They made our stays and visits very stimulating. ## 2. Picard–Fuchs system in the hyperelliptic case ### 2.1. Gelfand–Leray residue The derivative of an Abelian integral $`_{H=t}\omega `$ can be computed as the integral over the same curve of another 1-form $`\theta `$ called the Gelfand–Leray derivative (residue). More precisely, if a pair of polynomial 1-forms $`\omega ,\theta `$ satisfies the identity $`d\omega =dH\theta `$, then for any continuous family of cycles $`\delta (t)`$ on the level curves $`\{H=t\}`$ $$\frac{d}{dt}_{\delta (t)}\omega =_{\delta (t)}\theta ,\delta (t)\{H=t\}$$ (2.1) (the Gelfand–Leray formula). The identity remains true if $`\theta `$ is only meromorphic but has zero residues after restriction on each curve $`H=t`$. The identity between $`\omega `$, $`dH`$ and $`\theta `$ explains the standard notation $`\theta =d\omega /dH`$: to find $`\theta `$, one has to divide $`d\omega `$ by $`dH`$. In general this division is not possible in the class of polynomial 1-forms, but one can always divide $`d\omega `$ by $`dH`$ with remainder: the corresponding identity after integration will give a differential equation relating Abelian integrals with their derivatives. We illustrate this idea by deriving explicitly the Picard–Fuchs system for hyperelliptic Hamiltonians. In the hyperelliptic case the outlined approach yields a complete and in some sense minimal (irredundant) system that could be in principle derived by a number of different ways, e.g., as in . Moreover, using the explicit nature of Euclid’s algorithm of division of univariate polynomials, one can produce explicit upper bounds for the magnitude of the coefficients of the resulting equations, that are difficult (if possible at all) to obtain applying methods from . The constructions from this section serve as a paradigm for further exposition §3. ### 2.2. Division by polynomial ideals and 1-forms Let $`q_1,q_2[x,y]`$ be a pair of polynomials generating the ideal $`q_1,q_2[x,y]`$ that has a finite codimension $`\mu `$. By definition, this means that there exist $`\mu `$ polynomials $`r_1,\mathrm{},r_\mu [x,y]`$ (the remainders) such that any polynomial $`f[x,y]`$ admits representation $`v=q_1u_2q_2u_1+_1^\mu \lambda _ir_i`$ with polynomials $`u_1,u_2[x,y]`$ and constants $`\lambda _i`$. It is convenient to interpret this identity as a division formula for polynomial 2-forms: any polynomial 2-form $`\mathrm{\Omega }=f(x,y)dxdy`$ can be divided by the given 1-form $`\xi =q_1dx+q_2dy`$ with the “incomplete ratio” $`\eta =u_1dx+u_2dy`$ and the remainder that is a linear combination of the 2-forms $`\mathrm{\Omega }_i=r_idxdy`$, $$\mathrm{\Omega }=\xi \eta +\underset{i=1}{\overset{\mu }{}}\lambda _i\mathrm{\Omega }_i.$$ Denoting by $`\mathrm{\Lambda }^k`$, $`k=0,1,2`$, the modules (over the ring $`[x,y]`$) of polynomial $`k`$-forms on $`^2`$, we say that the tuple of 2-forms $`\{\mathrm{\Omega }_i\}_1^\mu `$ generates the quotient $`\mathrm{\Lambda }^2/\xi \mathrm{\Lambda }^1`$. The gradient ideal $`H_x,H_y`$ has a finite codimension provided that $`H`$ has only isolated critical points. In this case we will usually apply the division formula to a differential $`\mathrm{\Omega }=d\omega `$ of a polynomial 1-form and write the generators explicitly as $`\mathrm{\Omega }_i=d\omega _i`$ for appropriate polynomial primitives $`\omega _i\mathrm{\Lambda }^1`$: $$d\omega =dH\eta +\underset{i=1}{\overset{\mu }{}}\lambda _id\omega _i.$$ (2.2) This means that the Gelfand–Leray derivative of the form $`\omega _1^\mu \lambda _i\omega _i`$ can be found in the class of polynomial 1-forms, $`\eta \mathrm{\Lambda }^1`$. ### 2.3. Derivation of the Picard–Fuchs system in the hyperelliptic case Throughout this section we assume that $`H(x,y)=\frac{1}{2}y^2+p(x)`$, where $`p[x]`$ is a monic polynomial of degree $`n+1`$ in one variable: $`p(x)=x^{n+1}+_{i=0}^{n1}c_ix^i`$, $`|c_i|=c`$. The gradient ideal and the corresponding quotient in this case can be easily computed: $$H_x,H_y=p^{}(x),y,[x,y]/H_x,H_y[x]/x^n\underset{k=1}{\overset{n}{}}x^{k1},$$ so that the quotient algebra is an algebra of truncated univariate polynomials of degree $`n1`$. This observation motivates the following computation. Denote by $`\omega _i=x^{i1}ydx`$, $`i=1,\mathrm{},n`$, the differential 1-forms whose derivatives $`d\omega _i=x^{i1}dxdy`$ generate $`\mathrm{\Lambda }^2/dH\mathrm{\Lambda }^1`$. Then $`Hd\omega _i`$ $`=\left(\frac{1}{2}y^2+p(x)\right)x^{i1}dxdy`$ $`=\left[\frac{1}{2}x^{i1}yH_y+\left(b_i(x)H_x+a_i(x)\right)\right]dxdy`$ $`=\left(\frac{1}{2}x^{i1}ydxb_i(x)dy\right)dH+a_i(x)dxdy`$ $`=\left[\frac{1}{2}\omega _i+{\displaystyle \underset{j=1}{\overset{n}{}}}b_{ij}\omega _j+d(yb_j(x))\right]dH+{\displaystyle \underset{j=1}{\overset{n}{}}}a_{ij}d\omega _j,`$ where we used the following identities: 1. division with remainder: the polynomial $`x^{i1}p(x)`$ of degree $`n+i`$ is divided out by $`p^{}(x)=H_x`$ as $$x^{i1}p(x)=b_i(x)p^{}(x)+a_i(x),\mathrm{deg}b_ii,\mathrm{deg}a_in,$$ (2.3) 2. the form $`b_i(x)dy`$ is represented as a linear combination of the basic forms modulo an exact term: $$b_i(x)dy=d(yb_i)b_i^{}(x)ydx=\underset{j=1}{\overset{i}{}}b_{ij}x^{j1}ydx+dF_i,$$ (2.4) since the degree of $`b_i^{}[x]`$ never exceeds $`i1`$; 3. the remainders $`a_i(x)dxdy`$ can be represented as linear combinations of $`d\omega _j`$: $$a_i(x)dxdy=\underset{j=1}{\overset{n}{}}a_{ij}x^{j1}dxdy=\underset{j=1}{\overset{n}{}}a_{ij}d\omega _j.$$ (2.5) Integrating over closed ovals of the level curves $`H=t`$ (so that the exact forms $`dF_i`$ disappear) and using the Gelfand–Leray formula (2.1), we conclude with the system of linear ordinary differential equations $$t\dot{I}_i\underset{j=1}{\overset{n}{}}a_{ij}\dot{I}_j=\frac{1}{2}I_i+\underset{j=1}{\overset{n}{}}b_{ij}I_j$$ (2.6) or, in the matrix form, $$(tA)\dot{I}=BI,I^n,A,B\mathrm{Mat}_{n\times n}(),$$ (2.7) where, obviously, $`I_j(t)=\omega _j`$ are the Abelian integrals and $`I=(I_1,\mathrm{},I_n)`$ the column vector. ###### Remark 1. The computation above does not depend on the choice of the cycle of integration, therefore the system of equations will remain valid if we replace the column vector $`I`$ by the period matrix $`X(t)`$ obtained by integrating all forms $`\omega _i`$ over all vanishing cycles $`\delta _j(t)`$, $`j=1,\mathrm{},n`$ (see ) on the hyperelliptic level curves. The matrices $`A,B`$ can be completely described using the division process (2.3). Let $`x_{}`$ be a critical point of $`p`$ and $`t_{}=p(x_{})`$ the corresponding critical value. Then (2.3) imply that the column vector $`(1,x_{},x_{}^2,\mathrm{},x_{}^{n1})^n`$ is the eigenvector of $`A`$ with the eigenvalue $`t_{}`$, which gives a complete description (eigenbasis and eigenvalues) of $`A`$. Entries of the matrix $`B`$ can be described similarly: $`b_{ij}=0`$ for $`j>i`$ because of the assertion about degrees of $`b_j(x)`$, so $`B`$ is triangular. The diagonal entries can be easily computed by looking at the leading terms: since $`p`$ is monic, $`b_i(x)=\frac{x^i}{n+1}+\mathrm{}`$, hence $`b_i^{}(x)=\frac{i}{n+1}x^{i1}+\mathrm{}`$. Finally, the diagonal entries $`\{\frac{i}{n+1}+\frac{1}{2}\}_{i=1}^n`$ form the spectrum of $`B`$. However, knowledge of the critical values of $`H`$ is not yet sufficient to produce an upper bound for the norms $`A,B`$, since the conjugacy by the Vandermonde matrix (whose columns are the above eigenvectors $`(1,x_j,x_j^2,\mathrm{},x_j^{n1})^T`$, $`j=1,\mathrm{},n`$) may increase arbitrarily the norm of the diagonal matrix $`\mathrm{diag}(t_1,\mathrm{},t_n)`$, where $`t_j`$ are all critical values of $`H`$ (or $`p`$, what is the same). On the contrary, a linear change in the space of 1-forms that makes $`A`$ diagonal, can increase in an uncontrollable way the norm of the matrix $`B`$, whose eigenbasis differs from the standard one by a triangular transformation. It is the explicit division procedure that allows to majorize the matrix norms. ### 2.4. Bounds for the matrix norms For a polynomial $`p[x]`$ let $`p`$ be the sum of absolute values of its coefficients (sometimes it is called the length of $`p`$). It has the advantage of being multiplicative, $`pqpq`$. ###### Proposition 1. If $`q=x^n+\mathrm{}[x]`$ is a monic polynomial with $`qx^n=c`$, then any other polynomial $`f[x]`$ of degree $`dn`$ can be divided with remainder, $$f(x)=b(x)q(x)+a(x),\mathrm{deg}an1,$$ (2.8) so that $$b+aKf,K=1+C+C^2+\mathrm{}+C^{dn},C=1+c=q.$$ (2.9) ###### Proof. The proof goes by direct inspection of the Euclid algorithm of univariate polynomial division. The assertion of the Proposition is trivial for $`q=x^n`$: in this case the string of coefficients of $`r`$ has to be split into two, and immediately we have the decomposition $`r=bx^n+a`$ with $`b+a=r`$. The general nonhomogeneous case is treated by induction. Suppose that the inequality (2.9) is valid for any polynomial $`\stackrel{~}{f}`$ of degree $`d1`$ (for $`d=n1`$ it is trivially satisfied by letting $`b=0`$ and $`a=\stackrel{~}{f}`$). Take a polynomial $`f`$ of degree $`d`$ and write the identity $`f=bx^n+a=bq+b(x^nq)+a=bq+\stackrel{~}{f}`$, where the polynomial $`\stackrel{~}{f}=a+b(x^nq)`$ is of degree $`d1`$ and has the norm explicitly bounded: $`\stackrel{~}{f}cb+a(1+c)(b+a)Cf`$. By the induction assumption, $`\stackrel{~}{f}`$ can be divided, $`\stackrel{~}{f}=\stackrel{~}{b}q+\stackrel{~}{a}`$, with the norms satisfying the inequality (2.9). Collecting everything together, we have $`f=(b+\stackrel{~}{b})q+\stackrel{~}{a}`$ and $`b+\stackrel{~}{b}+\stackrel{~}{a}f+Cf(1+C+\mathrm{}+C^{d1n})f(1+\mathrm{}+C^{dn})`$. ∎ As a corollary to this Proposition and the explicit procedure of the division, we obtain upper bounds for norms of the matrices $`A,B`$. Recall that we use the $`\mathrm{}^1`$-norm on the “space of columns”, so the norm of a matrix $`A=\left(a_{ij}\right)_{i,j=1}^n`$ is $$A=\underset{j=1,\mathrm{},n}{\mathrm{max}}\underset{i=1}{\overset{n}{}}|a_{ij}|$$ (2.10) ###### Theorem 1. Suppose that $`p(x)=x^{n+1}+_{i=0}^{n1}c_ix^i`$ is a monic polynomial of degree $`n+1`$ and the non-principal part of $`p`$ is explicitly bounded: $`_{i=0}^{n1}|c_i|c`$. Then the entries of the matrices $`A,B`$ determining the Picard–Fuchs system (2.7) are explicitly bounded: $$A+Bn^2(1+C+\mathrm{}+C^{n+1}),C=1+c=p.$$ (2.11) ###### Proof. The derivative $`p^{}(x)`$ is not monic, but the leading coefficient is explicitly known: $`p^{}(x)=(n+1)(x^n+\mathrm{})`$, with the non-principal part denoted by the dots bounded by $`c`$ in the sense of the norm. Applying Proposition 1 to $`q=p^{}/(n+1)`$, we see that any polynomial can be divided by $`p^{}`$ and the same inequalities (2.9) would hold (since $`n+11`$). Thus we have $`b_i+a_iKx^{i1}p=KC`$, where $`K=1+C+\mathrm{}+C^n`$, then obviously $`b_j^{}nb_j`$ and finally for the sum of matrix elements $`A,B`$ occurring in the $`i`$th line, we produce an upper bound $`\frac{1}{2}+_j|b_{ij}|+_j|a_{ij}|nC(1+C+\mathrm{}+C^n)+\frac{1}{2}n(1+C+\mathrm{}+C^{n+1})`$. Clearly, this means that every entry of these matrices is majorized by the same expression and therefore for the matrix $`\mathrm{}^1`$-norms on $`^n`$ we have the required estimate. ∎ ### 2.5. Digression: doubly hyperelliptic Hamiltonians The algorithm suggested above, works with only minor modifications for doubly hyperelliptic Hamiltonians having the form $`H(x,y)=p(x)+q(y)`$ (the hyperelliptic case corresponds to $`q(y)=\frac{1}{2}y^2`$). Assume that $`n+1=\mathrm{deg}_xp`$, $`m+1=\mathrm{deg}_yq`$ (there is no reason to require that $`n=m`$). In this case the quotient algebra by the gradient ideal is generated by $`nm`$ monomials $`x^iy^j`$, $`0in1`$, $`0jm1`$. We claim that any collection of monomial primitives $`\omega _{ij}`$ to the monomial 2-forms $`d\omega _{ij}=x^iy^jdxdy`$ satisfies a system of $`nm`$ equations having the same form (2.7) though a different size. Indeed, $$Hd\omega _{ij}=p(x)x^idxy^jdyq(y)y^jdyx^idx.$$ Dividing the 1-form $`p(x)x^idx`$ with remainder by the 1-form $`dp(x)`$, we express the former as $`b_i(x)dp(x)+a_i(x)dx`$ with $`\mathrm{deg}b_ii+1n+1`$, $`\mathrm{deg}a_in1`$ and multiply the result by $`y^jdy`$. The second term can similarly be rewritten involving the representation $`q(y)y^j=b_j^{}(y)dq+a_j^{}(y)dy`$. Putting everything together, we conclude that $$Hd\omega _{ij}=[dp(x)b_i(x)y^jdydq(y)b_j^{}(y)x^idx]+[a_i(x)y^jx^ia_j^{}(y)]dxdy$$ Since $`dH=dp(x)+dq(y)`$, we see that the first bracket is actually the wedge product $`dH\eta _{ij}`$, where $`\eta _{ij}=b_i(x)y^jdy+b_j^{}(y)x^idx`$ is a polynomial 1-form whose differential $$d\eta _{ij}=\left(\frac{b_i}{x}y^j\frac{b_j^{}}{y}x^i\right)dxdy$$ has the coefficient of degree $`i`$ in $`x`$ and $`j`$ in $`y`$ and hence can be expanded as a linear combination of the forms $`\omega _{ij}`$ modulo an exact form. The second bracket, being a 2-form with coefficient of degrees $`n1`$ in $`x`$ and $`m1`$ in $`y`$, is a linear combination of the forms $`d\omega _{ij}`$. Thus we have the equations $$\begin{array}{c}Hd\omega _{ij}=dH\left(\underset{k,l=0}{}B_{ij,kl}\omega _{kl}+dF_{ij}\right)+\underset{k,l}{}A_{ij,kl}d\omega _{kl},\\ i,k=0,\mathrm{},n1,j,l=0,\mathrm{},m1.\end{array}$$ Rescaling $`x`$ and $`y`$ by appropriate factors independently, we can assume that the polynomials $`p(x)`$ and $`q(y)`$ are both monic. Then all divisions will be bounded provided that the norms $`p`$ and $`q`$ are explicitly bounded, and in a way completely similar to the arguments from §2.4, we can derive upper bounds for the matrix coefficients $`A_{ij,kl},B_{ij,kl}`$. Thus the case of doubly hyperelliptic Hamiltonians does not differ much from the ordinary hyperelliptic case, at least as far as the Picard–Fuchs systems for Abelian integrals are concerned. ### 2.6. Discussion The Picard–Fuchs system written in the form (2.7) for a generic hyperelliptic Hamiltonian (with the potential $`p(x)`$ being a Morse function on $``$), is a system remarkable for several instances: * it possesses only Fuchsian singularities (simple poles) both at all finite singularities $`t=t_j`$, $`j=1,\mathrm{},n`$, and at infinity; * it has no apparent singularities: all points $`t_j`$ are ramification points for the fundamental system of solutions $`X(t)`$ that is obtained by integrating all forms $`\omega _i`$ over all vanishing cycles $`\delta _j(t)`$ (the period matrix); * it is minimal in the sense that analytic continuations of any column of the period matrix $`X(t)`$ along all closed loops span the entire space $`^n`$. * its coefficients can be explicitly bounded in terms of $`H`$. (All these observations equally apply to doubly hyperelliptic Hamiltonians.) In the next section we generalize this result for arbitrary bivariate Hamiltonians. It will be impossible to preserve all properties, and we shall concentrate on the derivation of redundant system, eventually exhibiting apparent singularities, but all of them (including that at infinity) Fuchsian and with explicitly bounded coefficients. ## 3. Derivation of the redundant Picard–Fuchs system ### 3.1. Notations and conventions Recall that $`\mathrm{\Lambda }^k`$ denote the spaces of polynomial $`k`$-forms on $`^2`$ for $`k=0,1,2`$. They will be always equipped with the $`\mathrm{}^1`$-norms: the norm of a form is always equal to the sum of absolute values of all its coefficients. This norm behaves naturally with respect to the (wedge) product: for any two forms $`\eta \mathrm{\Lambda }^k`$, $`\theta \mathrm{\Lambda }^l`$, $`0k+l2`$, we always have $`\eta \theta \eta \theta `$. It is also convenient to grade the spaces of polynomial forms so that the degree of a $`k`$-form is the maximal degree of its (polynomial) coefficients plus $`k`$. Under this convention the exterior derivation is degree-preserving: $`\mathrm{deg}d\theta =\mathrm{deg}\theta `$ (unless $`d\theta =0`$). An easy computation shows that $`d\theta \mathrm{deg}\theta \theta `$ for any 0- and 1-form $`\theta `$. On several occasions the finite-dimensional linear space of $`k`$-forms of degree $`d`$ will be denoted by $`\mathrm{\Lambda }_d^k`$. If $`\omega \mathrm{\Lambda }^1`$ is a polynomial 1-form and $`H\mathrm{\Lambda }^0`$, then by $`d\omega /dH`$ is always denoted the Gelfand–Leray derivative (2.1), while by $`\frac{d\omega }{dxdy}`$ we denote the polynomial coefficient of the 2-form $`d\omega `$. The space $`\mathrm{\Lambda }^2`$ sometimes will be identified with $`\mathrm{\Lambda }^0[x,y]`$, the submodule $`dH\mathrm{\Lambda }^1`$ with the gradient ideal $`H_x,H_y[x,y]`$, and the local algebra as a linear space over $``$ with the quotient $`\mathrm{\Lambda }^2/dH\mathrm{\Lambda }^1`$. ### 3.2. Normalizing conditions and quasimonic Hamiltonians In the ring $`[x]`$ of univariate polynomials division by the principal ideal $`p`$ is a linear operator whose norm can be controlled in terms of $`p`$ provided that the leading term of $`p`$ is bounded from below, in particular when the polynomial is monic. The definition below introduces a generalization of this condition for ideals in the ring $`[x,y]`$ of bivariate polynomials. Recall that two homogeneous polynomials $`a,b[x,y]`$ of the same degree $`n`$ have no common linear factors if and only if their resultant is nonzero and hence the Sylvester matrix is invertible. In this case an arbitrary homogeneous polynomial $`f`$ of degree $`2n1`$ can be represented as $`f=au+bv`$ with appropriate (uniquely defined) homogeneous polynomials $`u,v`$ of degree $`n1`$ each. ###### Definition 2. A pair of homogeneous polynomials $`a,b[x,y]`$ of degree $`n`$ is said to be normalized if the linear operator $`(u,v)au+bv`$ restricted on the subspace of pairs of homogeneous polynomials of degree $`n1`$, has the inverse of the unit norm, in other words, if any homogeneous polynomial $`f`$ of degree $`2n1`$ can be represented as $`f=au+bv`$ with an explicit control over norms of the homogeneous “ratios” $`u,v`$ of degree $`n1`$: $$f=au+bv,u+vf.$$ (3.1) ###### Definition 3. A homogeneous polynomial 1-form $`\eta =adx+bdy`$ of degree $`n+1`$ is normalized if its coefficients $`a,b[x,y]`$ form a normalized pair. For nonhomogeneous objects we impose normalizing conditions on their principal homogeneous part. ###### Definition 4. A polynomial 1-form $`\xi \mathrm{\Lambda }^1`$ of degree $`n`$ is normalized at infinity, if its principal homogeneous part $`\widehat{\xi }`$ is normalized. A Hamiltonian $`H(x,y)[x,y]`$ of degree $`n+1`$ is said to be normalized at infinity or quasimonic, if $`dH`$ is normalized at infinity in the sense of the previous definition. ###### Remark 2. To be normalized at infinity has nothing to do with the $`\mathrm{}^1`$-norm of a form or Hamiltonian. We will mostly use the term “quasimonic”. ### 3.3. Balanced Hamiltonians In order to simplify the calculations below, we impose additional normalizing condition on $`H`$ meaning that the non-principal (low degree) terms are not dominating the principal part. ###### Definition 5. A Hamiltonian $`H[x,y]`$ will be called balanced, if it is quasimonic (the principal homogeneous part $`\widehat{H}`$ is normalized) and $`H\widehat{H}1`$. For a balanced Hamiltonian, its differential $`dH`$ is a 1-form that is (by definition) normalized at infinity and differs from its principal homogeneous part $`d\widehat{H}`$ by the form of degree $`n`$ and $`dHd\widehat{H}n`$. The two conditions, normalization at infinity and that of balance between principal and non-principal parts, can be obtained simultaneously by suitable affine transformations. If the Hamiltonian $`H`$ is regular at infinity, then after a suitable choice of $`\lambda `$ one can make any of the two polynomials, $`\lambda H(x,y)`$ or $`H(\lambda x,\lambda y)`$ being normalized at infinity (the same refers to 1-forms). Furthermore, if $`H`$ is already quasimonic, one can always choose a suitable $`\lambda `$ so that $`\lambda ^{n+1}H(\lambda ^1x,\lambda ^1y)`$ will be balanced while remaining quasimonic. ### 3.4. Lemma on bounded division Division by a balanced 1-form is a linear operator whose norm can be easily controlled. Let $`\xi \mathrm{\Lambda }^1`$ be a polynomial 1-form of degree $`n+1`$ normalized at infinity, with the principal homogeneous part denoted by $`\widehat{\xi }`$. ###### Lemma 1. Any polynomial 2-form $`\mathrm{\Omega }\mathrm{\Lambda }^2`$ can be divided with remainder by $`\xi `$, $$\mathrm{\Omega }=\xi \eta +\mathrm{\Theta },$$ (3.2) where the remainder $`\mathrm{\Theta }\mathrm{\Lambda }^2`$ is a 2-form of degree $`2n`$ and the “incomplete ratio” $`\eta \mathrm{\Lambda }^1`$ is a 1-form of degree $`\mathrm{deg}\mathrm{\Omega }\mathrm{deg}\xi `$. The decomposition (3.2) is in general non-unique. However, one can always find $`\eta `$ and $`\mathrm{\Theta }`$ so that if $`\xi \widehat{\xi }=c`$, then $$\eta +\mathrm{\Theta }K\mathrm{\Omega },K=(1+C+\mathrm{}+C^{d2n}),$$ (3.3) where $`C=c+1`$ and $`d=\mathrm{deg}\mathrm{\Omega }`$. ###### Proof. The proof reproduces almost literally the division algorithm for univariate polynomials, see Proposition 1. 1. For a homogeneous form $`\mathrm{\Omega }=fdxdy`$ of degree $`2n+1`$ the divisibility $`\mathrm{\Omega }=\widehat{\xi }\eta `$ by the homogeneous form $`\widehat{\xi }=adx+bdy`$ is the same as the representation (3.1) (recall that our convention concerning the degrees of the form means that in this case $`\mathrm{deg}f=2n1`$). From the normalization condition it follows then that $`\eta =u+vf=\mathrm{\Omega }`$ simply by definition. 2. Writing the division identities for all monomial forms of degree $`2n+1`$, multiplying them by arbitrary monomials and adding results we see then that any polynomial $`2`$-form $`\mathrm{\Omega }`$ containing no terms of degree $`2n`$ and less, can be divided by $`\widehat{\xi }`$ and the norm of the “ratio” $`\stackrel{~}{\eta }`$ does not exceed $`\mathrm{\Omega }`$. Finally, any form can be represented as the sum of a “remainder” $`\stackrel{~}{\mathrm{\Theta }}`$, the collection of terms of degree $`2n`$, and the higher terms divisible by $`\widehat{\xi }`$. All together this means that if $`\widehat{\xi }`$ is a homogeneous normalized 1-form of degree $`n+1`$, then any polynomial 2-form $`\mathrm{\Omega }`$ can be divided out as $$\mathrm{\Omega }=\widehat{\xi }\stackrel{~}{\eta }+\stackrel{~}{\mathrm{\Theta }},\stackrel{~}{\eta }+\stackrel{~}{\mathrm{\Theta }}\mathrm{\Omega },\mathrm{deg}\stackrel{~}{\eta }\mathrm{deg}\mathrm{\Omega }\mathrm{deg}\widehat{\xi }.$$ (3.4) 3. To divide by a nonhomogeneous form $`\xi `$ normalized at infinity, we first divide by its principal part $`\widehat{\xi }`$ as in (3.4). Then $$\mathrm{\Omega }=\xi \stackrel{~}{\eta }+(\widehat{\xi }\xi )\stackrel{~}{\eta }+\stackrel{~}{\mathrm{\Theta }}=\xi \stackrel{~}{\eta }+\stackrel{~}{\mathrm{\Omega }}.$$ (3.5) It remains to notice that $`\stackrel{~}{\eta }\stackrel{~}{\eta }+\stackrel{~}{\theta }\mathrm{\Omega }`$ and $`\stackrel{~}{\mathrm{\Omega }}`$ is a new 2-form whose degree is strictly less than $`d=\mathrm{deg}\mathrm{\Omega }`$, provided that $`d>2n`$. Since the norm of $`\xi \widehat{\xi }`$ is explicitly bounded by $`c`$, we have $$\stackrel{~}{\mathrm{\Omega }}c\stackrel{~}{\eta }+\stackrel{~}{\mathrm{\Theta }}(1+c)(\stackrel{~}{\eta }+\stackrel{~}{\mathrm{\Theta }})C\mathrm{\Omega }.$$ We may now continue by induction, accumulating the divided parts $`\stackrel{~}{\eta }`$ and reducing the degrees of “incomplete remainders” $`\stackrel{~}{\mathrm{\Omega }}`$ until the latter become less or equal to $`2n`$. More accurately, we use the inductive assumption to divide out $`\stackrel{~}{\mathrm{\Omega }}=\xi \eta ^{}+\mathrm{\Theta }`$ with $`\eta ^{}+\mathrm{\Theta }\stackrel{~}{\mathrm{\Omega }}(1+C+\mathrm{}+C^{d12n})\mathrm{\Omega }(C+C^2+\mathrm{}+C^{d2n})`$ and put $`\eta =\stackrel{~}{\eta }+\eta ^{}`$ so that $`\mathrm{\Omega }=\xi \eta +\mathrm{\Theta }`$. Since $`\stackrel{~}{\eta }\mathrm{\Omega }`$, we have $`\eta +\mathrm{\Theta }\eta +\eta ^{}+\mathrm{\Theta }(1+C+\mathrm{}+C^{d2n})\mathrm{\Omega }`$. ∎ ###### Corollary 1. If $`H`$ is a balanced Hamiltonian of degree $`n+1`$, then any polynomial $`2`$-form $`\mathrm{\Omega }`$ of degree $`3n`$ can be divided by $`dH`$, $$\mathrm{\Omega }=dH\eta +\mathrm{\Theta },\eta +\mathrm{\Theta }(n+1)^{n+1}\mathrm{\Omega }.$$ (3.6) ###### Proof of the Corollary. It is sufficient to remark that for a balanced Hamiltonian the form $`dH`$ is normalized at infinity and the difference between $`dH`$ and its principal homogeneous part $`d\widehat{H}`$ is of norm $`n`$. ∎ ### 3.5. Derivation of the redundant Picard–Fuchs system Now we can write explicitly a system of first order linear differential equations for Abelian integrals, with coefficients explicitly bounded provided the Hamiltonian is balanced (i.e., its lower order terms do not dominate the principal homogeneous part). The reason why this system is called redundant, will be explained below. Consider $`\nu =n(2n1)`$ monomial 2-forms $`\mathrm{\Omega }_i`$ spanning $`\mathrm{\Lambda }_{2n}^2`$, and let $`\omega _i\mathrm{\Lambda }_{2n}^1`$ be their monomial primitives (arbitrary chosen), $`d\omega _i=\mathrm{\Omega }_i`$, with unit coefficients so that $`\omega _i=1`$ and $`d\omega _i2n`$. Then any 2-form of degree $`2n`$ can be represented as a linear combination of $`d\omega _i`$, hence any 1-form of degree $`2n`$ admits representation as a linear combination of $`\omega _i`$, $`i=1,\mathrm{},\nu `$ modulo an exact differential. ###### Theorem 2. Let $`H`$ be a balanced Hamiltonian of degree $`n+1`$. Then the column vector $`I=(I_1(t),\mathrm{},I_\nu (t))`$ of integrals of all monomial 1-forms $`\omega _i`$ of degree $`2n`$ over any cycle on the level curves $`\{H(x,y)=t\}`$ satisfies the system of linear ordinary differential equations $$(tA)\dot{I}=BI,I=I(t)^\nu ,A,B\mathrm{Mat}_{\nu \times \nu }().$$ (3.7) The norms of the constant matrices $`A,B`$ are explicitly bounded: $$A+B6n(n+1)^{n+1}.$$ (3.8) ###### Remark 3. We use here the norms of matrices (2.10), associated with $`\mathrm{}^1`$-norms on the spaces of polynomials, as defined in (2.10). ###### Remark 4. As was already mentioned, the assumption that $`H`$ is balanced, does not involve loss of generality, since any Hamiltonian regular at infinity can be balanced by appropriate affine transformation (see however the discussion below). ###### Proof of the Theorem. We start with a computation showing that the system can be indeed written in the form (3.7): this derivation will be later slightly modified to produce explicit bounds. For any $`i=1,\mathrm{},\nu `$ the 2-form $`Hd\omega _i`$ of degree $`n+1+2n`$ can be divided out with remainder by the form $`dH`$ (which is, by assumption, normalized at infinity): $$Hd\omega _i=dH\eta _i+\mathrm{\Theta }_i,\mathrm{deg}\eta _i\mathrm{deg}d\omega _i2n,\mathrm{deg}\mathrm{\Theta }_i2n.$$ (3.9) Since 2-forms $`d\omega _i`$ span the whole space of 2-forms of degree $`2n`$, every $`d\eta _i`$ and $`\mathrm{\Theta }_i`$ are linear combinations of $`d\omega _j`$: $$d\eta _i=\underset{j=1}{\overset{\nu }{}}b_{ij}d\omega _j,\mathrm{\Theta }_i=\underset{j=1}{\overset{\nu }{}}a_{ij}d\omega _j,$$ with appropriate complex coefficients $`a_{ij},b_{ij}`$ forming two $`\nu \times \nu `$-matrices $`A,B`$ respectively, and certain polynomials $`F_j[x,y]`$. The first identity implies that $`\eta _i=_jb_{ij}\omega _j+dF_i`$ for suitable polynomials $`F_i`$. Integrating over cycles on the level curves $`\{H=t\}`$ and using the Gelfand–Leray formula for derivatives, we conclude that $$t\dot{I}_i=\underset{j}{}b_{ij}I_j+\underset{j}{}a_{ij}\dot{I}_j,i,j=1,\mathrm{},\nu ,$$ which is equivalent to the matrix form (3.7) claimed above. In order to place the upper bounds on the matrix norms $`A`$ and $`B`$, we can use the bounded division lemma, but additional efforts are required. Indeed, the normalization at infinity does not imply any upper bound on the norm of the principal part $`\widehat{H}`$, so the norm of the left hand side in (3.9) is apriori unbounded and does not allow for application of Lemma 1. To construct a system satisfying the inequalities (3.8), we decompose $`H`$ into the principal part $`\widehat{H}`$ and the collection of lower terms $`h=H\widehat{H}`$ and treat two parts, $`\widehat{H}d\omega `$ and $`hd\omega _i`$ separately. Let $`\rho \mathrm{\Lambda }_2^1`$ be the 1-form $`xdyydx`$ with $`\rho =2`$. Then by the Euler identity, $$(n+1)\widehat{H}dxdy=d\widehat{H}\rho ,$$ (3.10) and therefore $$\widehat{H}d\omega _i=\frac{d\omega _i}{dxdy}\widehat{H}dxdy=\frac{d\omega _i}{(n+1)dxdy}d\widehat{H}\rho =(dHdh)\eta _i^{},$$ (3.11) where $`\eta _i^{}d\omega _i\rho /(n+1)22n/(n+1)4`$. Now the term $`Hd\omega _i`$ can be explicitly expanded as $$Hd\omega _i=\widehat{H}d\omega _i+hd\omega _i=dH\eta _i^{}dh\eta _i^{}+hd\omega _i=dH\eta _i^{}+\mathrm{\Omega }_i^{},$$ where $`\mathrm{\Omega }_i^{}\eta ^{}dh+hd\omega _i4n+2n=6n`$ and $`\mathrm{deg}\mathrm{\Omega }_i^{}3n`$. Applying Corollary 1, we write $$\mathrm{\Omega }_i^{}=dH\eta _i^{\prime \prime }+\mathrm{\Theta }_i$$ with $`\eta _i^{\prime \prime }+\mathrm{\Theta }_i6n(n+1)^{n+1}`$ which together with the previous bounds for $`\eta _i^{}`$ would imply the inequality $$\eta _i+\mathrm{\Theta }_i6(n+1)^{n+2}.$$ (3.12) for the identities (3.9) Since all forms $`\omega _i,d\omega _i`$ are monomial with norms $`1`$, expanding $`\eta _i`$ and $`\mathrm{\Theta }_i`$ leads to coefficients satisfying the conditions $$\underset{j=1}{\overset{\mu }{}}|a_{ij}|\theta _i,\underset{j=1}{\overset{\mu }{}}|b_{ij}|\mathrm{\Theta }_i,$$ which gives the required bounds on $`A`$ and $`B`$. ∎ In order to incorporate the case of quasimonic but not balanced Hamiltonians, we derive an obvious corollary. ###### Corollary 2. If $`H`$ is quasimonic and the difference between $`H`$ and its principal homogeneous part $`\widehat{H}`$ is explicitly bounded, $$H\widehat{H}c,$$ (3.13) then one can choose the monomial forms so that the system (3.7) for their integrals involves the matrices $`A,B`$ satisfying the inequality $$A+B6(n+1)^{n+2}c^{n+1}.$$ (3.14) ###### Proof. It is sufficient to make a transformation replacing the initial Hamiltonian $`H(x,y)`$ by $`c^{(n+1)}H(cx,cy)`$. This will make $`H`$ balanced and the main theorem applicable. Notice that such transformation implies the change of time (the independent variable) $`tc^{(n+1)}`$ for the resulting system (3.7). With respect to the original variable the system (3.7) will take the form with the same matrix $`B`$, and $`A`$ multiplied by $`c^{n+1}`$. ∎ ###### Remark 5. Note that the system in the non-balanced case is written for forms $`\omega _i`$ in general not satisfying the condition $`\omega _i=1`$, as was the case with balanced Hamiltonians: the linear rescaling $`(x,y)(cx,cy)`$ results in a diagonal transformation that is in general non-scalar on the linear space of differential forms. ### 3.6. Abelian integrals of higher degrees The system of differential equations (3.7) holds for integrals of the basic monomial forms $`\omega _i`$ generating all polynomial differential 1-forms of degree $`2n`$. To write an analogous system for integrals of 1-forms of higher degrees, one can use the fact that the integrals $`\omega _i`$ generate the space of all Abelian integrals as a free $`[t]`$-module, provided that $`H`$ is Morse and regular at infinity, see . More precisely, if $`\mathrm{deg}\omega =d`$, then for any cycle $`\delta (t)`$ on the level curve $`\{H=t\}`$ one can represent $$_{\delta (t)}\omega =\underset{i=1}{\overset{\nu }{}}p_i(t)_{\delta (t)}\omega _i,p_i[t],(n+1)\mathrm{deg}p_i+\mathrm{deg}\omega _i\mathrm{deg}\omega ,$$ (3.15) (in fact, it is even sufficient to take any $`\mu =n^2`$ forms $`\omega _i`$ whose differentials span $`\mathrm{\Lambda }^2/dH\mathrm{\Lambda }^1`$). Thus the linear span of all functions $`t^kI_j(t)`$, $`j=1,\mathrm{},\nu `$, $`0km=d/(n+1)`$, contains all Abelian integrals of forms of degree $`d`$. The generators $`\{t^kI_j(t)\}_{0km}^{1jn}`$ of this system satisfy a block upper triangular system of linear first order differential equations obtained by derivation of (3.7): $$(tA)\frac{d}{dt}(t^kI)=Bt^kI+k(tA)t^{k1}I,k=1,\mathrm{},m.$$ (3.16) This system can be written in the matrix form involving two constant $`(m+1)\nu \times (m+1)\nu `$-matrices exactly as (3.7) and the entries of these matrices will be explicitly bounded, though this time the bounds and the size of the system will depend explicitly on $`d`$. Nevertheless this allows to treat integrals of forms of arbitrary fixed degree $`d`$ exactly as integrals of the basic forms. ### 3.7. Properties of the redundant Picard–Fuchs system Directly from the form in which the system (3.7) was obtained, it follows that it has singular points at all critical values $`t=t_j`$ of the Hamiltonian; the eigenvector corresponding to the eigenvalue $`t_j`$ has coordinates $`\frac{d\omega _i}{dxdy}(x_j,y_j)`$, $`i=1,\mathrm{},\nu `$. However, in general (since $`\mu <\nu `$) these eigenvalues do not exhaust the spectrum of $`A`$. The other eigenvalues of $`A`$ actually depend on the division with remainder, that is non-unique because the forms $`\omega _i`$ are linear dependent in $`\mathrm{\Lambda }^2/dH\mathrm{\Lambda }^1`$. Thus no invariant meaning can be associated with this part of the spectrum, and the corresponding singularities are apparent for the Abelian integrals (though other solutions can well have singularities at these “redundant” points). However, this freedom can be used to guarantee that all these singularities can be made Fuchsian, by slightly perturbing the matrices. ###### Proposition 2. If $`H`$ is a Morse function on $`^2`$, then the system (3.7) can be constructed so that the matrix $`A`$ has a simple spectrum while satisfying the same inequalities as before. ###### Corollary 3. The redundant system (3.7) can be always constructed having only Fuchsian singularities on the Riemann sphere. ###### Proof of the Corollary. The Fuchsian condition at infinity is satisfied automatically, as the matrix function $`(tA)^1B`$ has a simple pole at $`\tau =0`$ in the chart $`\tau =1/t`$. The inverse $`(tA)^1`$ can be obtained by dividing the adjugate matrix (a matrix polynomial of degree $`\nu 1`$ in $`t`$) by the determinant of $`tA`$, i.e., by the characteristic polynomial of $`A`$. As the latter has only simple roots by Proposition 2, all poles of $`(tA)^1B`$ at finite points are simple. ∎ ###### Proof of the Proposition. Assume that the enumeration of the forms $`\omega _i`$ is arranged so that the first $`\mu `$ of them constitute a basis in $`\mathrm{\Lambda }^2/dH\mathrm{\Lambda }^1`$. The procedure of division of the forms $`Hd\omega _i`$ by $`dH`$ can be altered to produce a unique answer, if we require that the remainder is always a linear combination of only the first $`\mu `$ forms. Moreover, instead of dividing the forms $`Hd\omega _i`$ with $`\mu +1i\nu `$, we will divide the forms $`(H\lambda _i)d\omega _i`$ with arbitrarily chosen constants $`\lambda _i`$, $`i=\mu +1,\mathrm{},\nu `$: $`Hd\omega _i`$ $`{\displaystyle \underset{j=1}{\overset{\mu }{}}}a_{ij}d\omega _jdH\mathrm{\Lambda }^1,i=1,\mathrm{},\mu ,`$ $`(H\lambda _i)d\omega _i`$ $`{\displaystyle \underset{j=1}{\overset{\mu }{}}}a_{ij}d\omega _jdH\mathrm{\Lambda }^1,i=\mu +1,\mathrm{},\nu .`$ After division organized in such a way, the matrix $`A`$ of the system (3.7) obtained after expanding the incomplete fractions, will have block lower-triangular form. The upper-left block of size $`\mu \times \mu `$ has as before the eigenvalues $`t_1,\mathrm{},t_\mu `$, while the lower-right block of size $`(\nu \mu )\times (\nu \mu )`$ is diagonal with $`\lambda _i`$ being the diagonal entries. Note that in this alternative derivation we lost control over the magnitude of the coefficients of remainders and incomplete ratios. Thus for the same column vector of Abelian integrals we have constructed two essentially different systems of the same form (3.7) but with different pairs $`(A,B)`$ of $`\nu \times \nu `$-matrices (the first bounded in the norm, the second with a predefined spectrum). By linearity, any linear homotopy between the two systems will also admit all Abelian integrals as solutions. Consider such a homotopy parameterized by $`s[0,1]`$. The eigenvalues of the matrix $`A`$ do depend algebraically on the parameter $`s`$. For $`s=1`$ they are equal to the critical values $`t_1,\mathrm{},t_\mu `$ of $`H`$ and arbitrarily prescribed values $`\lambda _{\mu +1},\mathrm{},\lambda _\nu `$. Since $`t_it_j`$ and $`\lambda _i`$ can be also chosen different from all $`t_j`$ and from each other, the eigenvalues are simple for $`s=1`$ and hence they remain pairwise different for almost all values of $`s`$, in particular, for arbitrarily small positive $`s`$ when the system is arbitrarily close to the first system (of explicitly bounded norm). Perturbing in that way achieves simplicity of the spectrum of the matrix $`A`$ while changing the norms of $`A,B`$ arbitrarily small. ∎ ## 4. Zeros of Abelian integrals away from the singular locus and related problems on critical values of polynomials ### 4.1. Heuristic considerations As was already noted in the introduction, the main reason why so much emphasis was put on explicit upper bounds for the coefficients of the system (3.7), was applications to the tangential Hilbert 16th problem. We explain in this section how the information accumulated so far can be used to place an effective upper bound for the number of zeros of Abelian integrals on a positive distance from the critical (ramification) locus and also on the total number of zeros of any branch for Hamiltonians whose critical values are distant from each other. However, the group of affine transformations acts naturally on the space of Abelian integrals in a quasihomogeneous manner, and to be geometrically sound, upper bounds for the number of zeros should be compatible with this symmetry. In particular, the above mentioned “positive distance to the critical locus” (resp., “distance between the critical values”) should be invariant by affine rescaling of Hamiltonians. Besides intrinsic considerations, the need for the bounds invariant by this action is motivated by the future study of zeros of Abelian integrals near singularities (cf. with ). ¿From the analytic point of view, the problem is in the choice of normalization on the variety of Hamiltonians regular at infinity. The geometric invariance requires this normalization to be imposed in terms of geometry of configurations of the critical values of the Hamiltonians. On the other hand, the assertion of the theorem on zeros of Abelian integrals, derived from the explicit form of the system (3.7), uses pre-normalization in terms of the coefficients of the Hamiltonian, more precisely, the $`\mathrm{}^1`$-norms of its nonhomogeneity (the difference between $`H`$ and its principal homogeneous part). Thus in a natural way the problem on equivalence of the two normalizing conditions arises. It can be shown relatively easily that a quasimonic polynomial whose non-principal part is bounded from above (in the sense of the norm), has all critical values inside a disk of known radius shrinking to a point as the non-principal part tends to zero (Proposition 4 below). One might hope that a converse statement is also true: if all critical values of a Hamiltonian $`H`$ come very close to each other, then (eventually after appropriate translations in the preimage and the image) $`H`$ differs from its principal homogeneous part $`\widehat{H}`$ by a small polynomial. This fact indeed holds true for univariate (and hence hyperelliptic) polynomials, where we were able to produce explicit inequalities between the diameter of the critical locus $`\mathrm{diam}\mathrm{\Sigma }=\mathrm{max}_{i,j=1,\mathrm{},\mu }|t_it_j|`$ and the nonhomogeneity $`H\widehat{H}`$, see Theorem 6 and Corollary 6. Yet for the truly bivariate polynomials the problem turned out to be considerably harder, and the best we were able to do is to show that for any fixed principal part $`\widehat{H}`$ the above two normalizations are equivalent, but as different linear factors of $`\widehat{H}`$ approach each other, the equivalence explodes. The current section explains the above arguments in more details and introduces the problem on relationships between the spread of critical values and an effective nonhomogeneoty of polynomials. Known partial results in this sense are collected in the next section. ### 4.2. Meandering theorem and upper bounds for zeros of Abelian integrals A (scalar) linear ordinary differential equation with explicitly bounded coefficients admits an explicit upper bound for the number of isolated (real or complex) zeros of all its solutions, see . The system of equations (3.7) can be reduced to one linear equation of degree $`\nu ^2`$ with rational in $`t`$ coefficients in such a way that any linear combination $`u(t)=_{i=1}^\nu c_iI_i(t)`$ of the integrals $`I_i(t)`$ with constant coefficients $`c_1,\mathrm{},c_\nu `$, will be a solution to this equation: it is sufficient to find a linear dependence between any fundamental $`\nu \times \nu `$-matrix $`X(t)`$ and its derivatives up to order $`\nu ^21`$ over the field $`(t)`$ of rational functions. Unfortunately, this procedure does not allow to place any bound on the magnitude of coefficients of the resulting equation. Instead, in \[15, Appendix B\] we described an algorithm of derivation of another linear equation of much higher order, whose coefficients are polynomially depending on the coefficients of the initial system (3.7). This algorithm is explicit, so that all degrees and coefficients admit explicit upper bounds. As a result, the system (3.7) is reduced to a Fuchsian linear differential equation of the form $$\begin{array}{c}\mathrm{\Delta }^{\mathrm{}}(t)u^{(\mathrm{})}+h_\mathrm{}1(t)\mathrm{\Delta }^\mathrm{}1(t)u^{(\mathrm{}1)}+\mathrm{}+h_1(t)\mathrm{\Delta }(t)u^{}+h_0(t)u=0,\end{array}$$ (4.1) where $`\mathrm{\Delta }(t)=(tt_1)\mathrm{}(tt_\nu )`$ is the characteristic polynomial of the matrix $`A`$ and all polynomial coefficients $`h_i(t)[t]`$, $`i=0,\mathrm{},\mathrm{}1`$, have degrees $`\mathrm{deg}h_i`$ and heights $`h_i`$ explicitly bounded by elementary functions of $`n`$. It is important to note here that the bounds, though completely explicit, are enormously excessive, being towers (iterated exponents) of height $`4`$. The coefficients of the equation (4.1) are explicitly bounded from above on the complement to sublevel sets $`\{|\mathrm{\Delta }(t)|\epsilon \}`$ for every given positive $`\epsilon >0`$. At the roots of $`\mathrm{\Delta }`$ (eigenvalues of $`A`$) the equation (4.1) has Fuchsian singularities, but the eigenvalues of $`A`$ that are not critical values of $`H`$, are apparent singularities for all linear combinations of the Abelian integrals $`I_j`$ (see §2). Recall that $`\mathrm{\Sigma }`$ is the critical locus (collection of all critical values) of the Hamiltonian $`H`$. Let $`R`$ be a finite positive number and $`K_R\mathrm{\Sigma }`$ the set obtained by cutting the set $$\{t:j=1,\mathrm{},\mu |tt_j|>1/R,|t|<R\}$$ (4.2) along no more than $`\mu `$ line segments to produce a simply connected compact “on the distance $`1/R`$ from both $`\mathrm{\Sigma }`$ and infinity”. Applying a general theorem on oscillations of solutions of linear equations with bounded coefficients , we arrive to the following theorem. ###### Theorem 3 (see ). Let $`H`$ be a balanced Hamiltonian of degree $`n+1`$ and $`K_R`$ a compact on distance $`1/R`$ from the critical locus of $`H`$ in the sense of (4.2). Then the number of zeros inside $`K_R`$ of any Abelian integral of a form of degree $`d`$ does not exceed $`(2+R)^N`$, where $`N=N(n,d)`$ is a certain elementary function depending only on $`n`$ and $`d`$. The function $`N(n,d)`$ can be estimated from above by a tower of four stories (iterated exponent) and certainly gives a very excessive bound. Yet we would like to remark that this is absolutely explicit bound, involving no undefined constants. ###### Remark 6. The necessity of cutting in the definition of $`K_R`$ is due to the fact that Abelian integrals are multivalued and a choice of branch should be specified each time when zeros are counted. The coefficients of the equation (4.1) blow up as $`t\mathrm{\Sigma }`$, so no upper bound for zeros can be derived from the general theorem . However, if the singularity $`t_i`$ is apparent and distant from all other points, say, at least by $`1`$, then one can place an upper bound on the coefficients of (4.1) on the boundary of the disk $`\{|tt_i|=\frac{1}{2}\}`$ and then by \[25, Corollary 2.7\] the variation of argument of any solution along the boundary can be explicitly bounded and by the argument principle, this would imply an upper bound for the number of zeros also inside the disk, where the coefficients are very large. It turns out that a similar construction can be also carried out when $`t_i`$ is a true (non-apparent) singularity, provided that it is of Fuchsian type and the spectrum of the monodromy operator is on the unit circle. Suppose that a function $`u(t)`$ analytic in the punctured disk $`\{0<|tt_i|1\}`$ admits a finite representation $`u(t)=_{\lambda ,k}f_{k,\lambda }(t)(tt_i)^\lambda \mathrm{ln}^k(tt_i)`$ with coefficients $`f_{k,\lambda }`$ analytic in the closed disk $`\{|tt_i|1\}`$, involving only real exponents $`\lambda `$. If this function satisfies a linear ordinary differential equation (with a Fuchsian singularity at $`t=t_i`$) whose coefficients are explicitly bounded on the boundary circumference of this disk, then it is proved in \[25, Theorem 4.1\] that any branch of $`u`$ admits an upper bound for the number of zeros in this disk in terms of the magnitude of the coefficients on the boundary and the order of the equation (the first result of this type was proved in ). The assumption on the spectrum always holds for Abelian integrals, since the above exponents $`\lambda `$ are always rational (in particular, equal to $`1`$ for a Morse critical value). Thus the above result (together with the bounded meandering principle) can be applied to the tangential Hilbert problem provided that all critical values of the Hamiltonian are at least $`1`$-distant from each other. An arbitrary Morse Hamiltonian one can rescaled to such form, yet the number $`\mathrm{min}_{ij}|t_it_j|`$ will enter then into the expression for the bound. By analogy with the previous result, denote by $`K_{\mathrm{}}`$ a simply connected open set obtained by slitting $`\mathrm{\Sigma }`$ along rays connecting critical values with infinity. ###### Theorem 4. Let $`H`$ be a balanced Hamiltonian of degree $`n+1`$, whose critical values $`t_1,\mathrm{},t_\mu `$ satisfy for some positive $`R<\mathrm{}`$ the condition $$|t_it_j|1/R,|t_i|Rij.$$ Then the number of zeros inside $`K_{\mathrm{}}`$ of any Abelian integral of a form of degree $`d`$ does not exceed $`(2+R)^N^{}`$, where $`N^{}=N^{}(n,d)`$ is a certain elementary function depending only on $`n`$ and $`d`$. ###### Sketch of the proof. Multiplying the Hamiltonian by $`R`$ and applying the bounded meandering principle to the Picard–Fuchs system (3.7), we construct a scalar linear equation of a very large order, satisfied by all Abelian integrals, so that its coefficients are explicitly bounded on distance $`1`$ from the critical locus by an expression polynomial in $`R`$ as above. To count zeros of Abelian integral inside the set $`K_{\frac{1}{2}}`$, one can use Theorem 3. The remaining part $`K_{\mathrm{}}K_{\frac{1}{2}}`$ consists of disjoint disks of radius $`1/2`$ centered at the critical values $`t_i`$ and slit along radii. Theorem 4.1 from applies to ever such disk and gives an upper bound for the number of zeros in these disks, thus completing the proof. ∎ Note the difference between two apparently similar results: Theorem 3 gives a uniform upper bound for the number of zeros in a certain domain (depending on the Hamiltonian, but always nonvoid for Morse Hamiltonians regular at infinity). On the contrary, Theorem 4 formally solves the tangential Hilbert problem for all Morse Hamiltonians (giving an upper bound for the number of all zeros, wherever they occur), but the bound is not uniform and explodes when the Hamiltonian approaches the boundary of the set of Morse polynomials regular at infinity. ### 4.3. Affine group action and equivariant problem on zeros of Abelian integrals Consider the affine complex space of Hamiltonians $`=\mathrm{\Lambda }_{n+1}^0`$ and the space of 1-forms $`=\mathrm{\Lambda }_d^1`$ of a given degree $`d`$. The Abelian integrals are multivalued functions on $`\left((\times )𝚺\right)\times `$, where $`𝚺`$ is the global discriminant, $$𝚺\times ,𝚺=\{(t,H):\text{ }t\text{ is a critical value of }H\}.$$ The group $`G_2`$ of affine transformations of $`^2`$ and the group $`G_1`$ of affine transformations of $`^1`$ act naturally on $`\times `$, $$(H,t)\stackrel{g_2,g_1}{}(g_1Hg_2,g_1t),$$ leaving $`𝚺`$ invariant. The problem of counting zeros of Abelian integrals should be also formulated for subsets in $`(\times )𝚺`$ that are invariant by this action. To achieve this equivariant formulation, we follow the ideology of normal forms and choose a convenient representative from each orbit of the group action. To factorize by the action of $`G_1`$, we notice that any point set $`t_1,\mathrm{},t_\mu `$ not reducible to one point, can be put by a suitable affine transformation (or, what is equivalent, by the choice of a chart) to a configuration satisfying two conditions, $$t_1+\mathrm{}+t_\mu =0,\underset{t=1,\mathrm{},\mu }{\mathrm{max}}|t_j|=1,$$ (4.3) and such transformation is determined uniquely modulo rotation of $``$, preserving the Euclidean metric on $`^2`$. Any set $`𝒦`$ in $`(\times )𝚺`$ invariant by the $`G_1`$-action, leaves its trace on the $`t`$-plane as a subset $`K`$ disjoint from the points $`\mathrm{\Sigma }=\{t_j\}_1^\mu `$ and the distance from $`K`$ to $`\mathrm{\Sigma }`$ measured in this privileged chart, is the natural equivariant distance between $`𝒦`$ and $`𝚺`$. We arrive thus to the following equivariant formulation of the problem on zeros of Abelian integrals, restricted in the sense that it concerns only zeros distant from singularities (this terminology was recently suggested by Yu. Ilyashenko). ###### Problem 2 (Equivariant restricted tangential Hilbert 16th problem). Let $`H`$ be a Hamiltonian of degree $`n+1`$ regular at infinity, whose critical values $`t_1,\mathrm{},t_\mu `$, $`\mu =n^2`$, satisfy the normalizing conditions (4.3). For any finite $`R>0`$ it is required to place an upper bound for the number of isolated zeros of Abelian integrals $`_{H=t}\omega `$ of any form of degree $`d`$ in the sets $`K_R`$ as in (4.2). The bound should depend only on $`n,d`$ and $`R`$. ### 4.4. From Theorem 3 to Equivariant problem In order to derive from Theorem 3 a solution to the equivariant problem, one should try to find in the orbit of the $`G_2`$-action on $``$ a Hamiltonian as close to be balanced as possible. Indeed, if for some affine transformation $`gG_2`$ the Hamiltonian $`\stackrel{~}{H}=Hg`$ is already balanced, then integrals of any form $`\omega `$ over any level curve $`H=t`$ are equal to integrals of the form $`g^{}\omega `$ over the curve $`\stackrel{~}{H}=t`$ (by the simple change of variables in the integral). But as $`g:^2^2`$ is an affine map, the form $`g^{}\omega `$ is again a polynomial 1-form of the same degree as $`\omega `$, while the new Hamiltonian $`\stackrel{~}{H}`$ is balanced. Hence Theorem 3 can be applied to produce the upper bound for the number of zeros exactly in the form we need to solve the equivariant problem: the result will be automatically a bound polynomial in $`R`$ with the exponent depending only on $`d`$ and $`n`$. In fact, it is sufficient to find in the $`G_2`$-orbit of $`H`$ a Hamiltonian $`\stackrel{~}{H}`$ that would be quasimonic and whose difference from its principal homogeneous part $`\widehat{H}`$ would be of norm explicitly bounded in terms of $`n`$. Indeed, if $`\stackrel{~}{H}`$ is such a polynomial and $`\stackrel{~}{H}\widehat{H}\tau =\tau (n)`$, then the transformation $$\stackrel{~}{H}(x,y)H^{}(x,y)=\tau ^{(n+1)}\stackrel{~}{H}(\tau x,\tau y)$$ (4.4) will preserve the principal homogeneous part $`\widehat{H}`$ while dividing all other terms by appropriate positive powers of $`\tau `$ so that in any case $`H^{}\widehat{H}1`$. This means that $`H^{}`$ is balanced and Theorem 3 can be applied and will give a bound on zeros $`1/R`$-distant from the critical locus of $`H^{}`$ in terms of $`R,n,d`$ as required. The transformation (4.4) does not preserve the normalizing conditions (4.3), but the conclusion of Theorem 3 can be rescaled to produce an upper bound on zeros $`1/\tau ^{n+1}R`$-distant from the (normalized) critical locus of $`H`$, by a suitable power of $`R\tau ^{n+1}`$, which will give a solution to the equivariant problem. Recall that the balance condition consists of the two parts: the (quasimonic) normalization of the principal homogeneous terms and the unit bound for the norm of all non-principal terms. The first part can be easily achieved by a suitable $`G_2`$-action. Indeed, replacing $`H(x,y)`$ by $`H(\tau x,\tau y)`$, one can effectively multiply the principal homogeneous part $`\widehat{H}`$ by $`\tau ^{n+1}`$ and thus achieve the required normalization. It will be convenient in the future not to change the principal part any more, once it was made quasimonic. This means that the only remaining degree of freedom to use is the group of translations of $`^2`$ (and rotations that do not affect norms). Summarizing this discussion, we see that in order to derive from Theorem 3 the equivariant restricted tangential Hilbert 16th problem (Problem 2), it would be sufficient to solve the following problem. ###### Definition 6. For a quasimonic polynomial $`H`$ with the principal part $`\widehat{H}`$ we call its effective nonhomogeneity the lower bound $$\varkappa (H)=\underset{TG_2}{inf}HT\widehat{H},T\text{ a translation of }^2.$$ (4.5) ###### Problem 3. Given a quasimonic Hamiltonian $`H`$ of degree $`n+1`$, whose critical values satisfy the normalizing conditions (4.3), place an upper bound for the effective nonhomogeneity $`\varkappa (H)`$. This and related problem, completely independent from all previous considerations, is discussed and partially solved in the next section. ## 5. Critical values of polynomials ### 5.1. Geometric consequences of quasimonicity The normalizing condition at infinity (for 1-forms and Hamiltonians) was introduced in purely algebraic terms as an inequality imposed on the principal homogeneous part of a 1-form (resp., Hamiltonian). However, one can provide a simple geometric meaning to this condition. Recall that if $`H`$ is regular at infinity, then its principal homogeneous part $`\widehat{H}`$ has an isolated critical point at the origin. This means that the gradient $`\widehat{H}`$ never vanishes outside the origin, and in particular its minimal (Hermitian) length on the boundary of the unit bidisk $`𝔹=\{|x|1,|y|1\}^2`$ is strictly positive. Because of the homogeneity, this is sufficient to place a lower bound on the length of $`\widehat{H}`$ everywhere on $`^2\{0\}`$. ###### Proposition 3. If $`\widehat{H}`$ is normalized (quasimonic), then everywhere on the boundary of the unit bidisk $`𝔹`$ the Hermitian length of $`\widehat{H}`$ is no smaller than $`1`$. ###### Proof. Consider the part $`𝔹_1`$ of the boundary $`𝔹`$ which is given by the inequalities $`|x|=1`$, $`|y|1`$ (the other part is treated similarly). The homogeneous polynomial $`x^{2n1}`$ can be represented as $`a\widehat{H}_x+b\widehat{H}_y`$ with $`a+b1`$. Restricting this on $`𝔹_1`$ we see that the Hermitian product of the gradient $`\widehat{H}=(\widehat{H}_x,\widehat{H}_y)`$ and the vector field $`V`$ with coordinates $`(\overline{a},\overline{b})`$ is everywhere equal to 1 in the absolute value. The Hermitian length of $`V`$ at any point of $`𝔹_1`$ can be easily majorized by $`\sqrt{|\overline{a}(x,y)|^2+|\overline{b}(x,y)|^2}`$ which is no greater than $`\sqrt{a^2+b^2}1`$ on $`𝔹_1=\{|x|=1,|y|1\}`$. But then by the Cauchy inequality, the length of $`\widehat{H}`$ cannot be smaller than $`1`$ on $`𝔹_1`$. ∎ ### 5.2. Almost-homogeneity implies close critical values We begin by showing that a quasimonic Hamiltonian whose non-principal part is bounded, admits an upper bound for the moduli of critical values. This solves the problem inverse to Problem 3. ###### Proposition 4. If $`H`$ is a quasimonic Hamiltonian of degree $`n+1`$ with the principal part $`\widehat{H}`$ and $`H\widehat{H}\frac{1}{n\sqrt{2}}`$, then the critical values of $`H`$ are all in the disk $`\{|t|3/n\}`$. ###### Proof. Denote $`H=\widehat{H}+h`$. The gradient of each monomial of degree $`n`$ has the Hermitian length bounded by $`n\sqrt{2}`$ on the unit bidisk $`𝔹`$. Thus if $`h<\frac{1}{n\sqrt{2}}`$, then $`h`$ has its length strictly bounded by $`1`$ everywhere in $`𝔹`$. By Proposition 3, the length of $`\widehat{H}`$ is at least $`1`$ everywhere on the boundary of $`𝔹`$, so by the topological index theorem, all $`\mu =n^2`$ critical points of $`H`$ must be be inside $`𝔹`$. Note that a quasimonic principal part $`\widehat{H}`$ admits no apriori upper bound on $`𝔹`$, however, the critical values of $`H=\widehat{H}+h`$ can be explicitly majorized. Indeed, at any critical point $`(x_{},y_{})`$, the gradient of $`H`$ vanishes so $`\widehat{H}(x_{},y_{})=h(x_{},y_{})`$. By the Euler identity, $`(n+1)|\widehat{H}(x_{},y_{})|=|(x_{},y_{})\widehat{H}(x_{},y_{})|=|(x_{},y_{})h(x_{},y_{})|\sqrt{2}`$, since the Hermitian length of $`h(x,y)`$ is explicitly bounded by $`1`$ in $`𝔹`$. Finally, since $`|h(x,y)|h\frac{1}{n\sqrt{2}}`$ in $`𝔹`$, we conclude that $`|H(x_{},y_{})||\widehat{H}(x_{},y_{})|+|h(x_{},y_{})|\frac{\sqrt{2}}{n+1}+\frac{1}{n\sqrt{2}}3/n\sqrt{2}3/n`$. ∎ Thus when discussing the equivariant restricted Hilbert problem, only the other direction (Problem 3) is interesting. ### 5.3. Dual formulation, limit and existential problems Problem 3 can be reformulated in dual terms as follows. ###### Problem 4 (dual to Problem 3). Given a quasimonic Hamiltonian $`H`$ of effective nonhomogeneity $`\varkappa (H)=1`$, place a lower bound on the diameter of its critical values $$\mathrm{diam}\mathrm{\Sigma }=\underset{1ij\mu }{\mathrm{max}}|t_it_j|.$$ Having solved this problem, one can easily derive from it by the rescaling arguments as above a solution to Problem 3 and vice versa. The dual formulation of Problem 4 allows a limit version: one is required to show that if $`H`$ cannot be reduced to a homogeneous polynomial by a translation, i.e., $`\varkappa (H)>0`$, then $`\mathrm{diam}\mathrm{\Sigma }>0`$, i.e., not all critical values coincide. This limit problem can be settled. ###### Theorem 5. If a polynomial $`H(x,y)`$ regular at infinity has only one critical value (necessarily of multiplicity $`\mu =n^2`$), then by a suitable translations in the preimage and the image $`H`$ can be made homogeneous: $`H(x,y)=\widehat{H}(x+\alpha ,y+\beta )+\gamma `$, where $`\widehat{H}`$ is the principal homogeneous part of $`H`$. We postpone the proof of Theorem 5, deriving first as a corollary an existential solution of either of the two equivalent Problems 3 and 4. ###### Corollary 4. For a quasimonic Hamiltonian $`H=\widehat{H}+h`$ of degree $`n+1`$ there exist two positive finite constants, $`\alpha =\alpha (\widehat{H})`$ and $`\beta =\beta (\widehat{H})`$, depending only on the principal part $`\widehat{H}`$, such that the critical locus $`\mathrm{\Sigma }=\mathrm{\Sigma }(H)`$ and the effective non-homogeneity $`\varkappa (H)`$ are related as follows: $`\varkappa (H)1`$ $`\mathrm{\Sigma }\{|t|>\alpha \}\mathrm{},`$ (5.1) $`\mathrm{\Sigma }\{|t|1\}`$ $`\varkappa (H)\beta .`$ ###### Proof of the Corollary. Consider the affine space $`_n^{(n+1)(n+2)/2}`$ of polynomials of degree $`n`$ in two variables, and define two nonnegative functions on it, $$f(h)=\varkappa (\widehat{H}+h)=\underset{T^2}{inf}T^{}(\widehat{H}+h),g(h)=\underset{t\mathrm{\Sigma }(\widehat{H}+h)}{}|t|^2,$$ where $`T`$ ranges over all translations of the plane $`T^2`$ and $`\mathrm{\Sigma }(H)`$ is the collection of all critical values of the Hamiltonian $`H=\widehat{H}+h`$ with the fixed principal part $`\widehat{H}`$. Both functions, as one can easily see, are semilagebraic on $`^2^4`$. From Theorem 5 it follows that $`f(h)`$ must vanish if $`g(h)=0`$, i.e., that the zero locus of $`g`$ is contained in that of $`f`$. By the Łojasiewicz inequality, there exist two positive finite constants $`C,\rho >0`$, such that $$f(h)Cg^\rho (h),h_n.$$ From this inequality the assertion of the Corollary easily follows if we let $`\beta =Cn^\rho `$ and $`\alpha =(nC)^{1/2\rho }`$. Since $`C,\rho `$ depend only on the construction of $`f,g`$, that is, on $`\widehat{H}`$, the Corollary is proved. ∎ Unfortunately, the proof gives no means to compute explicitly the bounds $`\alpha `$ and $`\beta `$. Moreover, below we will show that they cannot be chosen uniformly over all quasimonic principal parts. ### 5.4. Parallel problems for univariate polynomials One can easily formulate analogs of all the above problems for univariate polynomials, in which case monic rather than quasimonic polynomials are to be considered. Note that the critical values of the hyperelliptic Hamiltonian $`H(x,y)=y^2+p(x)`$ coincide with that of the univariate potential $`p[x]`$, and also the effective nonhomogeneity (more accurately, non-quasihomogeneity) of $`H`$ coincides with that $`\varkappa (p)`$. Thus all results proved below, are valid not only for univariate polynomials, but also for hyperelliptic bivariate Hamiltonians. The limit problem for this case is fairly elementary. It was solved by A. Chademan as a step towards the existential solution of Problem 3 for univariate polynomials, see Corollary 5 below. ###### Proposition 5 (Chademan ). A complex polynomial that has only one critical value at $`t=0`$, is a translated monomial $`\alpha (xa)^{n+1}`$. ###### Proof. Assuming without loss of generality that the polynomial $`p(x)`$ is monic, we can always write the derivative $$p^{}(x)=(n+1)(xa_1)^{\nu _1}\mathrm{}(xa_k)^{\nu _k},$$ where $`a_1,\mathrm{},a_k`$ are geometrically distinct critical points and all $`\nu _k>0`$. For any $`j=1,\mathrm{},k`$ the polynomial $`p`$ can be expressed as the primitive of $`p^{}`$ integrated from $`a_j`$, $$p(x)=p(a_j)+_{a_j}^xp^{}(s)𝑑s=0+(xa_j)^{\nu _j+1}q_j(x),q_j[x],$$ in other words, $`p`$ is divisible by $`(xa_j)^{\nu _j+1}`$. As this holds for all points $`a_j`$, $`j=1,\mathrm{},k`$, hence $`\mathrm{deg}p\mathrm{deg}p^{}+k`$ and therefore only one $`\nu _j`$ can be different from zero. ∎ In the standard way (see the demonstration of Corollary 4 above) the following corollary can be derived. ###### Corollary 5 (A. Chademan ). If $`p(x)=x^{n+1}+p_{n1}x^{n1}+\mathrm{}+p_1x+p_0`$ is a monic polynomial of degree $`n+1`$ without the term $`x^n`$, and all complex critical values of $`p`$ lie in the unit disk $`\{|t|1\}`$, then $$|p_{n1}|+\mathrm{}+|p_1|+|p_0|C_n,$$ where $`C_n`$ is a constant depending only on $`n`$.∎ However, in the same way as before, the proof based on solution of the limit problem gives no possibility of effectively computing the constant $`C_n`$. We compute it using alternative approach. ### 5.5. Spread of roots vs. spread of critical values for univariate monic complex polynomials ###### Theorem 6. If all critical values $`\{t_1,\mathrm{},t_n\}`$ of a monic univariate polynomial $`p(x)=_{j=0}^n(xx_j)`$ are in the unit disk, then the diameter of the set of its roots is no greater than $`4e`$: $$\mathrm{\Sigma }\{|t|1\}j,k=0,\mathrm{},n|x_jx_k|4e.$$ ###### Proof. Consider the real-valued function $`f:`$, $`f(x)=|p(x)|`$. It is smooth outside the roots of the polynomial $`p`$. Moreover, its critical values (different from zero) coincide with $`|t_j|`$, as the critical points for $`f`$ and $`p`$ are the same. By the main principle of the Morse theory, all sublevel sets $`M_s=\{x:f(x)s\}`$ for $`0<s<\mathrm{}`$ of the function $`f`$ remain homeomorphic to each other until $`s`$ passes through a critical value of $`f`$. One can easily verify that the $`M_s`$ is simply connected for all large $`s`$ (it differs only slightly from the disk $`\{|t|s^{1/n+1}\}`$). Our assumption on the critical values guarantees that the set $`M_1=\{|p(x)|1\}`$ corresponding to $`s=1`$ is therefore also connected (though its shape can be very non-circular anymore). On the other hand, by the famous Cartan lemma for any positive $`\epsilon `$ one can delete from $``$ one or several disks with the sum of diameters less than $`\epsilon `$ so that on the complement the monic polynomial of degree $`n+1`$ satisfies a lower bound $`|p(x)|(\epsilon /4e)^{n+1}`$. This lemma implies that the set $`M_1`$ can be covered by one or several circular disks with the sum of diameters $`4e`$. But the set $`M_1`$ (like all sets $`M_s`$ with positive $`s`$) contains all roots of $`p`$, so if there are two roots $`x_i,x_j`$ on the distance more than $`4e`$, then the union of disks covering these two roots simultaneously, cannot be connected (it is sufficient to project all the disks on the line connecting these roots and reduce the assertion to one dimension). This contradiction proves the theorem. ∎ ###### Corollary 6. By a suitable translation $`p(x)p(x+a)`$ a monic polynomial $`p(x)=x^{n+1}+\mathrm{}`$ whose critical values are normalized by the conditions (4.3), can be reduced to the form $`p(x)=x^{n+1}+_{j=0}^np_jx^j`$ with $`_j|p_j|=x^{n+1}p8^{n+1}`$. ###### Proof. By Theorem 6, the roots of $`p`$ form a point set of diameter $`d4e`$ in the $`x`$-plane. Any such set can be covered by a regular hexagon with the opposite sides being at the distance $`d`$ . Shifting the origin at the center of this hexagon makes all roots $`x_j`$ satisfying the inequality $`|x_j|d/\sqrt{3}`$. A monic polynomial of degree $`n+1`$ with all roots inside the disk of radius $`r>0`$ has all its coefficients bounded by the respective coefficients of the polynomial $`(x+r)^{n+1}`$, by the Vieta formulas. For the latter polynomial the sum of (absolute values of) all coefficients is the value at $`x=1`$ (since all these coefficients are nonnegative). Putting everything together, we conclude that after shifting the origin at the center of the hexagon, $`p(x)(1+4e/\sqrt{3})^{n+1}8^{n+1}`$. ∎ ###### Remark 7. Simply shifting the origin to one of the roots makes all of them being in the circle of radius $`4e`$, which finally yields an upper bound $`p(1+4e)^{n+1}12^{n+1}`$ without referring to the claim on hexagonal cover. The assertion of Theorem 6 for real polynomials having only real critical points, can be proved in a completely different way. The following proposition gives an insight as to how accurate the bound established in Theorem 6 is. ###### Proposition 6. A monic real polynomial of degree $`n+1`$ with all critical points real and all critical values in the interval $`[1,1]`$, has all its real roots in some interval of the length $`4`$. ###### Proof. Between any two roots the polynomial satisfies the condition $`1p(x)1`$, since all critical values lie on that interval. Among monic polynomials of degree $`n+1`$ on the unit interval $`1x1`$ the smallest uniform upper bound $`c_n=2^{(n+1)}`$ is achieved for the Chebyshev polynomial $`T_n(x)=2^{(n+1)}\mathrm{cos}(n+1)\mathrm{arccos}x`$: for any other monic polynomial of this degree, the $`C^0`$-norm $`\mathrm{max}_{1x1}|p(x)|`$ will be greater or equal to $`c_n`$. Applying this assertion to the polynomial $`2^{n+1}p(x/2)`$ we conclude that the largest real interval on which the monic polynomial can satisfy the condition $`|p|1`$, is of length $`4`$ (twice the length of $`[1,1]`$). ∎ Thus Theorem 6 can be considered as generalizing (in some sense) the extremal property of the Chebyshev polynomials to the complex domain. ### 5.6. Demonstration of Theorem 5 The proof of Theorem 5 is an immediate corollary to the two following lemmas. ###### Lemma 2. A polynomial regular at infinity and having only one complex critical value, has a unique critical point. This lemma is in fact valid for polynomials of any number of variables. The second claim is dimension-specific. ###### Lemma 3. A bivariate polynomial regular at infinity and having a unique complex critical point at the origin, is homogeneous. ###### Proof of Lemma 2. Let $`H_\epsilon `$ be an analytic one-parameter perturbation of the polynomial $`H_0=H`$, such that for all $`\epsilon 0`$ the polynomial $`H_\epsilon `$ is Morse. Consider the monodromy group of the bundle $`H_\epsilon :^2^1`$ for an arbitrary small $`\epsilon `$. It is known that vanishing cycles form the basis of the homology of all fibers, each being a cyclic vector (i.e., all continuations of any vanishing cycle span the entire first homology of the typical fiber $`\{H_\epsilon =t\}`$. Suppose that there are at least two critical points $`a_1a_2^2`$ for $`H_0`$. Then for all sufficiently small $`\epsilon `$ the polynomial $`H_\epsilon `$ will have two disjoint groups of critical points with close critical values. Moreover, these groups of critical points are well apart (say, the distance between them is never smaller than half the distance between $`a_1`$ and $`a_2`$). But then the vanishing cycles “growing” from critical points not belonging to the same group, are also disjoint, therefore their intersection index must be zero. But then the Picard–Lefschetz formulas imply that the subspaces generated by each group of vanishing cycles, must be both invariant, which contradicts the fact that each group must consist of cyclic elements for the monodromy. ∎ ###### Proof of Lemma 3. Consider the one-parameter analytic (polynomial) homotopy between $`H`$ and its principal part, $`H_\epsilon (x,y)=\epsilon ^{n+1}H(\epsilon ^1x,\epsilon ^1y)`$. Then for $`\epsilon =0`$ $`H_0`$ coincides with the principal homogeneous part, while $`H_1=H`$. The germ of $`H_\epsilon `$ at the origin $`x=y=0`$ has a multiplicity $`\mu _\epsilon `$ (the Milnor number) that is equal to $`n^2`$ for any $`\epsilon `$. Indeed, by the Bézout theorem, the total number of critical points of $`H`$ counted with multiplicities in the projective plane $`P^2`$, is $`n^2`$; the condition of nondegeneracy at infinity implies that all of them are in the finite (affine) part $`^2`$. The uniqueness assumption means that all these $`n^2`$ points coincide at the origin. By the famous theorem due to D. T. Lê and C. P. Ramanujam , the topological type of an analytic germ is constant along the stratum $`\mu =\mathrm{const}`$, therefore the germs of $`H_0`$ and $`H_1`$ at the origin are topologically equivalent, in particular, the germs of analytic curves $`\{H_0=0\}`$ and $`\{H_1=0\}`$ in $`(^2,0)`$ are homeomorphic. But by the Zariski theorem , the order of a planar analytic curve (i.e., the order of the lowest order terms which occur in the Taylor expansion of the local equation defining this curve) is a topological invariant. For the curve $`H_0=0`$ this order is $`n+1`$, as the polynomial $`H_0`$ is homogeneous. But this means that the lowest order of terms that may occur in $`H_1`$, is also $`n+1`$, that is, $`H_0=H_1`$ and $`H`$ coincides in fact with its principal homogeneous part. ∎ ###### Remark 8. Consider the gradient vector field $`H`$. Its principal homogeneous part, $`\widehat{H}`$, is a homogeneous vector field on the plane that has an isolated singularity of multiplicity $`n^2`$ at the origin. Assertion of Lemma 3 means that adding any nontrivial lower order terms to $`H`$ would necessarily create singular points of the gradient vector field outside the origin, thus changing the multiplicity of what remains at the origin. However, this assertion about arbitrary (not necessarily gradient) polynomial vector fields is false, as the following example shows. ###### Example 1 (Lucy Moser–Jauslin). The nonhomogeneous vector field $$(x^3y^3+x)\frac{}{x}+(2x^3y^3+x)\frac{}{y}$$ has a unique singular point of the maximal multiplicity $`9`$ at the origin, and the principal homogeneous part has an isolated singularity. ### 5.7. Existential bounds cannot be uniform As was already noted, the proof of Corollary 4 gives no indication on how to compute the bounds $`\alpha (\widehat{H})`$ and $`\beta (\widehat{H})`$ for a given homogeneous part $`\widehat{H}`$. However, the folowing example shows that there cannot be the bound uniform over all principal parts: as some of the linear factors approach each other, the values of $`\beta `$ and $`\alpha ^1`$ may grow to infinity. ###### Example 2. The form $`\widehat{H}_a(x,y)=a\frac{x^{n+1}}{n+1}+\frac{y^{n+1}}{n+1}`$ is normalized for $`a1`$, as one can easily see by comparing the operator of division by $`d\widehat{H}=ax^n,y^n`$ on $`2n`$-forms with that by the ideal $`x^n,y^n`$. The polynomial $`H_a(x,y)=\widehat{H}_a(x,y)x`$ has critical points at $`y=0`$, $`x=1/\sqrt[n]{a}`$ (of multiplicity $`n`$ for every choice of branch of the root). The corresponding critical values all converge to zero asymptotically as $`a^{1/n}`$ as $`a\mathrm{}`$. On the other hand, the effective nonhomogeneity of the univariate polynomial $`p_a(x)=a\frac{x^{n+1}}{n+1}x`$ (and hence the value $`\varkappa (H_a)`$) remain bounded away from zero as $`a\mathrm{}`$. Indeed, if after shifting the polynomial $`p_a`$ by $`r=r(a)`$ the coefficient before $`x^n`$ goes to zero, then necessarily $`ar(a)0`$. On the other hand, the coefficient before the linear term is equal to $`1+ar(a)^n`$ and hence is bounded away from zero. Thus the bounds established in Corollary 4, cannot be made uniform over all homogeneous parts. Of course, the reason is that the space of quasimonic principal parts is not compact (e.g., the polynomials $`\widehat{H}_a`$ have no limit points as $`a\mathrm{}`$). In turn, this is related to the fact that some of the linear factors entering $`\widehat{H}_a`$, tend to each other (as points on the projective line $`P^1`$). ### 5.8. Discussion: atypical values and singular perturbations The phenomenon occurring in the above example, might be characteristic. When the Hamiltonian is not regular at infinity, the Abelian integrals may have ramification points that are not critical values of $`H`$. Such points, called atypical values, must necessarily be singular for any system of Picard–Fuchs equations, and are studied mostly by topological means. On the other hand, the fact that entries of the matrices $`A,B`$ may grow to infinity as the principal part of $`\widehat{H}`$ degenerates, means that the system (3.7) (written in the privileged chart to make the assertion equivariant) undergoes a singular perturbation (appearance of a large parameter in the right hand side that is equivalent to putting a small parameter before some of the higher order derivatives). Thus we see that “atypical singularities” in the Picard–Fuchs system can appear as a result of singular perturbation. The analytic approach based on studying division by $`dH`$ and arguments involving geometry of critical values, may be a complementary tool for the study of singularities “coming from infinity”.
warning/0001/astro-ph0001364.html
ar5iv
text
# New Techniques for Relating Dynamically Close Galaxy Pairs to Merger and Accretion Rates : Application to the SSRS2 Redshift Survey ## 1. INTRODUCTION Studies of galaxy evolution have revealed surprisingly recent changes in galaxy populations. Comparisons of present day galaxies with those at moderate ($`z0.5`$) and high ($`z3`$) redshift have uncovered trends which are often dramatic, and may trace galaxies to the time at which they were first assembled into recognizable entities. These discoveries have shed new light on the formation of galaxies, and have provided clues as to the nature of their evolution. At $`z<1`$, the picture that is emerging is one in which early type galaxies evolve slowly and passively, while late type galaxies become more numerous with increasing redshift (e.g., Lin et al. (1999)). At higher redshifts, deep surveys such as the Hubble Deep Field (Williams et al. (1996)) indicate an increase in the cosmic star formation rate out to $`z2`$ (e.g., Madau, Pozzetti, and Dickinson 1998). While considerable progress has been made in the observational description of galaxy evolution, important questions remain regarding the physical processes driving this evolution. Mechanisms that have been postulated include galaxy-galaxy mergers, luminosity-dependent luminosity evolution, and the existence of a new population of galaxies that has faded by the present epoch (see reviews by Koo & Kron (1992) and Ellis (1997)). In this study, we will investigate the relative importance of mergers in the evolution of field galaxies. Mergers transform the mass function of galaxies, marking a progression from small galaxies to larger ones. In addition, mergers can completely disrupt their constituent galaxies, changing gas-rich spiral galaxies into quiescent ellipticals (e.g., Toomre and Toomre 1972). During a collision, a merging system may also go through a dramatic transition, with the possible onset of triggered star formation and/or accretion onto a central black hole (see review by Barnes & Hernquist 1992). It is clear that mergers do occur, even during the relatively quiet present epoch. However, the frequency of these events, and the distribution of masses involved, has yet to be accurately established. This is true at both low and high redshift. Furthermore, while a number of attempts have been made, a secure measurement of evolution in the galaxy merger rate remains elusive, and a comparable measure of the accretion rate has yet to be attempted. In this study, we introduce a new approach for relating dynamically close galaxy pairs to merger and accretion rates. These new techniques yield robust measurements for disparate samples, thereby allowing meaningful comparisons of mergers at low and high redshift. In addition, these pair statistics can be adapted to a variety of redshift samples, and to studies of both major and minor mergers. We apply these techniques to a large sample of galaxies at low redshift (SSRS2), providing a much needed local benchmark for comparison with samples at higher redshift. In a forthcoming paper (Patton et al. 2000), we will apply these techniques to a large sample of galaxies at moderate redshift (CNOC2; $`0.1<z<0.6`$), yielding a secure estimate for the rate of evolution in the galaxy merger and accretion rates. An overview of earlier pair studies, and a discussion of their limitations and shortcomings, are given in the next section. The SSRS2 data are described in § 3. Section 4 discusses the connection between close pairs and the merger and accretion rates, while § 5 introduces new statistics for relating these quantities. Section 6 describes how these statistics can be applied to flux-limited surveys in a robust manner. A pair classification experiment is presented in § 7, giving empirical justification for our close pair criteria. Pair statistics are then computed for the SSRS2 survey in § 8, and the implications are discussed in § 9. Conclusions are given in the final section. Throughout this paper, we use a Hubble constant of $`H_0=100h`$ km s<sup>-1</sup> Mpc<sup>-1</sup>. We assume $`h`$=1 and $`q_0`$=0.1, unless stated otherwise. ## 2. BACKGROUND Every estimate of evolution in the merger and/or accretion rate begins with the definition of a merger statistic. Ideally, this statistic should be independent of selection effects such as optical contamination due to unrelated foreground/background galaxies, redshift incompleteness, redshift-dependent changes in minimum luminosity resulting from flux limits, contamination due to non-merging systems, $`k`$-corrections, and luminosity evolution. In addition, it should be straightforward to relate the statistic to the global galaxy population, and to measurements on larger scales. The statistic should then be applied to large, well-defined samples from low to high redshift, yielding secure estimates of how the merger and/or accretion rates vary with redshift. Within the past decade, there have been a number of attempts to estimate evolution in the galaxy merger rate using close pairs of galaxies (e.g., Zepf & Koo (1989), Burkey et al. (1994), Carlberg, Pritchet, & Infante (1994), Woods, Fahlman, & Richer (1995), Yee & Ellingson (1995), Patton et al. (1997), Le Fèvre et al. (1999)). The statistic that has been most commonly employed is the traditional pair fraction, which gives the fraction of galaxies with suitably close physical companions. This statistic is assumed to be proportional to the galaxy merger rate. The local (low-redshift) pair fraction was estimated by Patton et al. (1997), using a flux-limited ($`B14.5`$) sample of galaxies from the UGC catalog (Nilson (1973)). Using pairs with projected physical separations of less than 20 $`h^1`$ kpc, they estimated the local pair fraction to be $`4.3\pm 0.4\%`$. This result was shown to be consistent with the local pair fraction estimates of Carlberg et al. (1994) and Yee & Ellingson (1995), both of whom also used the UGC catalog. The pair fraction has been measured for samples of galaxies at moderate redshift ($`z0.5`$), yielding published estimates ranging from approximately 0% (Woods, Fahlman, & Richer (1995)) to 34% $`\pm `$ 9% (Burkey et al. (1994)). Evolution in the galaxy merger rate is often parameterized as $`(1+z)^m`$. Close pair studies have yielded a wide variety of results, spanning the range $`0<m<5`$. There are several reasons for the large spread in results. First, different methods have been used to relate the pair fraction to the merger rate. In addition, some estimates have been found to suffer from biases due to optical contamination or redshift completeness. After taking all of these effects into account, Patton et al. (1997) demonstrated that most results are broadly consistent with their estimate of $`m=2.8\pm 0.9`$, made using the largest redshift sample (545 galaxies) to date. While this convergence seems promising, all of these results have suffered from a number of very significant difficulties. The central (and most serious) problem has been the comparison between low and moderate redshift samples. Low-$`z`$ samples have been poorly defined, due to a lack of suitable redshift surveys. In addition, the pair fraction depends on both the clustering and mean density of galaxies. The latter is very sensitive to the limiting absolute magnitude of galaxies, leading to severe redshift-dependent biases when using flux-limited galaxy samples. These biases have not been taken into account in the computation of pair fractions, or in the comparison between samples at different redshifts. While these problems are the most serious, there are several other areas of concern. A lack of redshift information has meant dealing with optical contamination due to unrelated foreground and background galaxies. Moreover, while one can statistically correct for this contamination, it is still not possible to discern low velocity companions from those that are physically associated but unbound, unless additional redshift information is available. Finally, there is no direct connection between the pair fraction and the galaxy correlation function (CF) and luminosity function (LF), making the results more difficult to interpret. To address these issues, we have developed a novel approach to measuring pair statistics. We will introduce new statistics that overcome many of the afflictions of the traditional pair fraction. We will then apply these statistics to a large, well-defined sample of galaxies at low redshift. ## 3. DATA The Second Southern Sky Redshift Survey (da Costa et al. (1998); hereafter SSRS2) consists of 5426 galaxies with $`m_B15.5`$, in two regions spanning a total of 1.69 steradians in the southern celestial hemisphere. The first region, denoted SSRS2 South, has boundaries $`40{}_{}{}^{}\delta 2.5^{}`$ and $`b_{II}40^{}`$. The second region, SSRS2 North, is a more recent addition, and is bounded by $`\delta 0^{}`$ and $`b_{II}35^{}`$. Galaxies were selected primarily from the list of non-stellar objects in the Hubble Space Telescope Guide Star Catalog, with positions accurate to $``$ 1<sup>′′</sup> and photometry with an rms scatter of $``$ 0.3 magnitudes (Alonso et al. 1993, Alonso et al. 1994). Steps were taken to ensure that single galaxies were not mistakenly identified as close pairs, due to the presence of dust lanes, etc. (da Costa et al. (1998)). In addition, careful attention was paid to cases where a very close pair might be mistaken for a single galaxy. This was found to make a negligible contribution to the catalog as a whole ($`<0.1\%`$ of galaxies are affected). The effect on the pairs analysis in this paper is further reduced by imposing a minimum pair separation of 5 $`h^1`$ kpc (see § 7). The sample now includes redshifts for all galaxies brighter than $`m_B15.5`$. We correct all velocities to the local group barycenter using Equation 6 from Courteau and van den Bergh (1999). We restrict our analysis to the redshift range $`0.005z0.05`$. This eliminates nearby galaxies, for which recession velocities are dominated by peculiar velocities, giving poor distance estimates. We also avoid the sparsely sampled high redshift regime. This leaves us with a well-defined sample of 4852 galaxies. ## 4. THE GALAXY MERGER RATE AND ACCRETION RATE ### 4.1. Definitions The primary goal of earlier close pair studies has been to determine how the galaxy merger rate evolves with redshift. The merger rate affects the mass function of galaxies, and may also be connected to the cosmic star formation rate. Before attempting to measure the merger rate, it is important to begin with a clear definition of a merger and a merger rate. Here, we refer to mergers between two galaxies which are both above some minimum mass or luminosity. If this minimum corresponds roughly to a typical bright galaxy ($`L_{}`$), this criterion can be thought of as selecting so-called major mergers. We consider two merger rate definitions. First, it is of interest to determine the number of mergers that a typical galaxy will undergo per unit time. In this case, the relevant rate may be termed the galaxy merger rate (hereafter $`_{\mathrm{mg}}`$). A related quantity is the total number of mergers taking place per unit time per unit co-moving volume. We will refer to this as the volume merger rate (hereafter $`_{\mathrm{mg}_V}`$). Clearly, $`_{\mathrm{mg}_V}=n_0_{\mathrm{mg}}`$, where $`n_0`$ is the co-moving number density of galaxies. While both of these merger rates provide useful measures of galaxy interactions, they have their limitations. As one probes to faint luminosities, one will find an increasing number of faint companions; hence, the number of inferred mergers will increase in turn. For all realistic LFs, this statistic will become dominated by dwarf galaxies. In addition, it is of interest to determine how the mass of galaxies will change due to mergers. To address these issues, we will also investigate the rate at which mass is being accreted onto a typical galaxy. This quantity, the total mass accreted per galaxy per unit time, will be referred to as the galaxy accretion rate (hereafter $`_{\mathrm{ac}}`$). This is related to the rate of mass accretion per unit co-moving volume ($`_{\mathrm{ac}_V}`$) by $`_{\mathrm{ac}_V}=n_0_{\mathrm{ac}}`$. The mass (or luminosity) dependence of the accretion rate means that it will be dominated by relatively massive (or luminous) galaxies, with dwarfs playing a very minor role unless the mass function is very steep. ### 4.2. Observable Quantities In order to determine $`_{\mathrm{mg}}`$ observationally, one may begin by identifying systems which are destined to merge. By combining information about the number of these systems and the timescale on which they will undergo mergers, one can estimate an overall merger rate. Specifically, if one identifies $`N_m`$ ongoing mergers per galaxy, and if the average merging timescale for these systems is $`T_{\mathrm{mg}}`$, then $`_{\mathrm{mg}}=N_m/T_{\mathrm{mg}}`$. If the mass involved in these mergers (per galaxy) is $`M_m`$, then $`_{\mathrm{ac}}=M_m/T_{\mathrm{mg}}`$. In practice, direct measurement of these quantities is a daunting task. It is difficult to determine if a given system will merge; furthermore, estimating the merger timescale for individual systems is challenging with the limited information generally available. However, if one simply wishes to determine how the merger rate is changing with redshift, then the task is more manageable. If one has the same definition of a merger in all samples under consideration, then it is reasonable to assume that the merger timescale is the same for these samples. In this case, we are left with the task of measuring quantities which are directly proportional to the number or mass of mergers per galaxy or per unit co-moving volume. If one wishes to consider luminosity instead of mass, the relation between mass and luminosity must either be the same at all epochs, or understood well enough to correct for the differences. We have considered several quantities that fit this description. All involve the identification of close physical associations of galaxies. A “close companion” is defined as a neighbour which will merge within a relatively short period of time ($`T_{\mathrm{mg}}t_H`$), which allows an estimate of the instantaneous merger/accretion rate. If a galaxy is destined to undergo a merger in the very near future, it must have a companion close at hand. One might attempt to estimate the number of mergers taking place within a sample of galaxies. For example, a close pair of galaxies would be considered one merger, while a close triple would lead to two mergers, etc. Owing to the difficulty of determining with certainty which systems are undergoing mergers, we will not use this approach. One alternative is to estimate the number of galaxies with one or more close companions, otherwise known as the pair fraction. One drawback of this approach is that close triples or higher order N-tuples complicate the analysis, since they are related to higher orders of the correlation function. This also makes it difficult to correct for the flux-limited nature of most redshift surveys. As a result, we choose to steer clear of this method also. In this study, we choose instead to use the number and luminosity of close companions per galaxy. The number of close companions per galaxy, hereafter $`N_c`$, is similar in nature to the pair fraction. In fact, they are identical in a volume-limited sample with no triples or higher order N-tuples. However, $`N_c`$ will prove to be much more robust and versatile. We assume that $`N_c`$ is directly proportional to the number of mergers per galaxy, such that $`N_m=kN_c`$ ($`k`$ is a constant). This pairwise statistic is preferable to the number of mergers per galaxy or the fraction of galaxies in merging systems, in that it is related, in a direct and straightforward manner, to the galaxy two-point CF and the LF (see Section 5). We note that it is not necessary that there be a one-to-one correspondence between companions and mergers, as long as the correspondence is the same, on average, in all samples under consideration. Using this approach to estimate the number of mergers per galaxy, the merger rate is then given by $`_{\mathrm{mg}}=kN_c/T_{\mathrm{mg}}`$. The actual value of $`k`$ depends on the merging systems under consideration. If one identifies a pure set of galaxy pairs, each definitely undergoing a merger, then each pair, consisting of 2 companions, would lead to one merger, giving $`k`$=0.5. For a pair sample which includes some triples and perhaps higher order N-tuples, $`k<0.5`$. If the merging sample under investigation contains some systems which are not truly merging (for instance, close pairs with hyperbolic orbits), then $`k`$ will also be reduced. While $`k`$ clearly varies with the type of merging system used, the key is for $`k`$ to be the same for all samples under consideration. We take a similar approach with the accretion rate. We again use close companions, and in this case we simply add up the luminosity in companions, per galaxy ($`L_c`$). Defining the mean companion mass-to-light ratio as $`\mathrm{{\rm Y}}`$, it follows that $`M_m=\mathrm{{\rm Y}}L_c`$ and $`_{\mathrm{ac}}=\mathrm{{\rm Y}}L_c/T_{\mathrm{mg}}`$. When comparing different samples, any significant differences in $`\mathrm{{\rm Y}}`$ must be accounted for. ### 4.3. A Simple Model of Mass Function Evolution Due to Mergers In order to motivate further the need for merger rate measurements, and to set the stage for future work relating pair statistics to the mass and luminosity function, we develop a simple model which relates these important quantities. Suppose the galaxy mass function is given by $`\varphi (M,t)`$. This function gives the number density of galaxies of mass $`M`$ at time $`t`$, per unit mass. The model that follows can also be expressed in terms of luminosity or absolute magnitude, rather than mass. We begin by assuming that all changes in the mass function are due to mergers. While this is clearly simplistic, this model will serve to demonstrate the effects that various merger rates can have on the mass function. In order to relate the mass function to the observable luminosity function, we further assume that mergers do not induce star formation. Again, this is clearly an oversimplification; however, this simple case will still provide a useful lower limit on the relative contribution of mergers to LF evolution. Finally, we assume that merging is a binary process. Following Bahcall & Tremaine (1988), we consider how $`\varphi (M,t)`$ evolves as the universe ages from time $`t`$ to time $`t+\mathrm{\Delta }t`$. Each merger will remove two galaxies from the mass function, and produce one new galaxy. Let $`\mathrm{\Delta }\varphi (M,\mathrm{\Delta }t)_{\mathrm{out}}`$ represent the decrease in the mass function due to galaxies removed by mergers, while $`\mathrm{\Delta }\varphi (M,\mathrm{\Delta }t)_{\mathrm{in}}`$ gives the increase due to the remnants produced by these mergers. Evolution in the mass function can then be given by $$\varphi (M,t+\mathrm{\Delta }t)=\varphi (M,t)\mathrm{\Delta }\varphi (M,\mathrm{\Delta }t)_{\mathrm{out}}+\mathrm{\Delta }\varphi (M,\mathrm{\Delta }t)_{\mathrm{in}}.$$ (1) We model this function by considering all galaxy pairs, along with an expression for the merging likelihood of each. Let $`p(M,M^{},\mathrm{\Delta }t)`$ denote the probability that a galaxy of mass $`M`$ will merge with a galaxy of mass $`M^{}`$ in time interval $`\mathrm{\Delta }t`$. In order to estimate $`\mathrm{\Delta }\varphi (M,\mathrm{\Delta }t)_{\mathrm{out}}`$, we need to take all galaxies of mass $`M`$, and integrate over all companions, yielding $`\mathrm{\Delta }\varphi (M,\mathrm{\Delta }t)_{\mathrm{out}}=`$ $`{\displaystyle _0^{\mathrm{}}}\varphi (M,t)\varphi (M^{},t)p(M,M^{},\mathrm{\Delta }t)[h^1\mathrm{Mpc}]^3𝑑M^{}.`$ (2) We devise a comparable expression for $`\mathrm{\Delta }\varphi (M,\mathrm{\Delta }t)_{\mathrm{in}}`$ by integrating over all pairs with end$``$products of mass $`M`$. This is achieved by considering all pairs with component of mass $`MM^{}`$ and $`M^{}`$, such that $`\mathrm{\Delta }\varphi (M,\mathrm{\Delta }t)_{\mathrm{in}}={\displaystyle _0^{\mathrm{}}}\varphi (MM^{},t)\varphi (M^{},t)`$ $`\times p(MM^{},M^{},\mathrm{\Delta }t)[h^1\mathrm{Mpc}]^3dM^{}.`$ (3) We can also express Equation 1 in terms of the pair statistics outlined in Section 4.2. If one considers close companions of mass $`M^{}\mathrm{}`$ next to primary galaxies of mass $`M`$, the volume merger rate can be expressed as $`_{\mathrm{mg}_V}(M)`$, yielding $$\mathrm{\Delta }\varphi (M,\mathrm{\Delta }t)_{\mathrm{out}}=_{\mathrm{mg}_V}(M)\mathrm{\Delta }t.$$ (4) Similarly, if one defines a merger remnant statistic, $`_{\mathrm{mr}_V}(M)`$, to be the co-moving number density of merger remnants per unit time corresponding to these same mergers, then $$\mathrm{\Delta }\varphi (M,\mathrm{\Delta }t)_{\mathrm{in}}=_{\mathrm{mr}_V}(M)\mathrm{\Delta }t.$$ (5) Therefore, it is possible, in principle, to use pair statistics to measure the evolution in the mass or luminosity function due to mergers. However, current pair samples are too small to permit useful pair statistics for different mass combinations. In addition, present day observations of close pairs are not of sufficient detail to determine the proportion of pairs that will result in mergers (factor $`k`$ in previous section). Moreover, timescale estimates for these mergers are not known with any degree of certainty. Hence, useful observations of mass function evolution due to mergers will have to wait for improved pair samples and detailed estimates of merger timescales. ## 5. A NEW APPROACH TO MEASURING PAIR STATISTICS In this section, we outline the procedure for measuring the mean number ($`N_c`$) and luminosity ($`L_c`$) of close companions for a sample of galaxies with measured redshifts. We begin by defining these statistics in real space, demonstrating how they are related to the galaxy LF and CF. We then show how these statistics can be applied in redshift space. ### 5.1. Pair Statistics in Real Space In this study, we will measure pair statistics for a complete low-redshift sample of galaxies (SSRS2). However, we wish to make these statistics applicable to a wide variety of redshift samples. We would also like this method to be useful for studies of minor mergers, where one is interested in faint companions around bright galaxies. Moreover, these techniques should be adaptable to redshift samples with varying degrees of completeness (that is, with redshifts not necessarily available for every galaxy). Therefore, in the following analysis, we treat host galaxies and companions differently. Consider a primary sample of $`N_1`$ host galaxies with absolute magnitudes $`MM_1`$, lying in some volume $`V`$. Suppose this volume also contains a secondary sample of $`N_2`$ galaxies with $`MM_2`$. In the general case, the primary and secondary samples may have galaxies in common. This includes the special case in which the two samples are identical. If $`M_1M_2`$, this will tend to probe major mergers. If $`M_2`$ is chosen to be significantly fainter than $`M_1`$, this will allow for the study of minor mergers. We assume here that both samples are complete to the given absolute magnitude limits; in Section 6, we extend the analysis from volume-limited samples to those that are flux-limited. We wish to determine the mean number and luminosity of companions (in the secondary sample) for galaxies in the primary sample. In real space, we define a close companion to be one that lies at a true physical separation of $`rr^{\mathrm{max}}`$, where $`r^{\mathrm{max}}`$ is some appropriate maximum physical separation. To compute the observed mean number ($`N_c`$) and luminosity ($`L_c`$) of companions, we simply add up the number ($`N_{c_i}`$) and luminosity ($`L_{c_i}`$) of companions for each of the $`N_1`$ galaxies in the primary sample, and then compute the mean. Therefore, $$N_c(MM_2)=\underset{i=1}{\overset{N_1}{}}N_{c_i}/N_1$$ (6) and $$L_c(MM_2)=\underset{i=1}{\overset{N_1}{}}L_{c_i}/N_1.$$ (7) We can also estimate what these statistics should be, given detailed knowledge of the galaxy two-point CF $`\xi `$ and the LF $`\varphi (M)`$. This is necessary if one wishes to relate these pair statistics to measurements on larger scales. Consider a galaxy in the primary sample with absolute magnitude $`M_i`$ at redshift $`z_i`$. We would first like to estimate the number and luminosity of companions lying in a shell at physical distance (proper co-ordinates) $`rr_{ij}r+dr`$ from this primary galaxy. To make this estimate, we need to know the mean density of galaxies (related to the LF), and the expected overdensity in the volume of interest (given by the CF). The mean physical number density of galaxies at redshift $`z_i`$ in the secondary sample, with absolute magnitudes $`MM_jM+dM`$, is given by $$n_2(z_i,M)dM=(1+z_i)^3\varphi (M,z_i)dM,$$ (8) where $`\varphi (M,z_i)`$ is the differential galaxy LF, which specifies the co-moving number density of galaxies at redshift $`z_i`$, in units of $`h^3\mathrm{Mpc}^3\mathrm{mag}^1`$. The actual density of objects in the region of interest is determined by multiplying the mean density by (1+$`\xi `$), where $`\xi `$ is the overdensity given by the two-point CF (Peebles (1980)). In general, $`\xi `$ depends on the pair separation $`r`$, the mean redshift $`z_i`$, the absolute magnitude of each galaxy ($`M_i`$, $`M_j`$), and the orbits involved, specified by components parallel ($`v_{}`$) and perpendicular ($`v_{}`$) to the line of sight. It follows that the mean number of companions with $`MM_jM+dM`$ and $`rr_{ij}r+dr`$ is given by $`N_{c_i}(z_i,M_i,M,r,v_{},v_{})dMdr=`$ $`n_2(z_i,M)dM[1+\xi (r,z_i,M_i,M,v_{},v_{})]4\pi r^2dr.`$ (9) We must now integrate this expression for all companions with $`MM_2`$ and $`rr^{\mathrm{max}}`$. Integration over the LF yields $$n_2(z_i,MM_2)=(1+z_i)^3_{\mathrm{}}^{M_2}\varphi (M,z_i)𝑑M.$$ (10) Integration over the CF is non-trivial, because of the complex nature of $`\xi (r,z_i,M_i,M,v_{},v_{})`$. With redshift samples that are currently available, it is not possible to measure this dependence accurately for the systems of interest. Hence, we must make three important assumptions at this stage. First, we assume that $`\xi `$ is independent of luminosity. Later in the paper, we demonstrate empirically that this is a reasonable assumption, provided one selects a sample with appropriate ranges in absolute magnitude (see Section 6.1). Secondly, we assume that the distribution of velocities is isotropic. If one averages over a reasonable number of pairs, this is bound to be true, and therefore $`\xi `$ is independent of $`v_{}`$ and $`v_{}`$. Finally, we assume that the form of the CF, as measured on large scales, can be extrapolated to the small scales of interest here. This assumption applies only to the method of relating pairs to large scale measures, and not to the actual measurement of pair statistics. It is now straightforward to integrate Equation 5.1. The mean number of companions with $`MM_2`$ and $`rr^{\mathrm{max}}`$ for a primary galaxy at redshift $`z_i`$ is given by $`N_{c_i}(z_i,MM_2,rr^{\mathrm{max}})=`$ $`n_2(z_i,MM_2){\displaystyle _0^{r^{\mathrm{max}}}}[1+\xi (r,z_i)]4\pi r^2𝑑r.`$ (11) We derive an analogous expression for the mean luminosity in companions. The integrated luminosity density is given by $$j_2(z_i,MM_2)=(1+z_i)^3_{\mathrm{}}^{M_2}\varphi (M,z_i)L(M)𝑑M,$$ (12) where $$L(M)=10^{0.4(MM_{})}L_{}.$$ (13) Therefore, $`L_{c_i}(z_i,MM_2,rr^{\mathrm{max}})=`$ $`j_2(z_i,MM_2){\displaystyle _0^{r^{\mathrm{max}}}}[1+\xi (r,z_i)]4\pi r^2𝑑r.`$ (14) Given measurements of the CF on large scales, it is then straightforward to integrate these equations to arrive at predicted values of $`N_{c_i}`$ and $`L_{c_i}`$. It is important to note that these statistics are directly dependent on $`M_2`$, which affects the mean density of galaxies in the secondary sample. This is different from statistics such as the correlation function, which are independent of density. Hence, this serves as a reminder that we must exercise caution when choosing our samples, to ensure that differences in the pair statistics (and hence in the merger and accretion rates) are not simply due to apparent density differences resulting from selection effects. In addition, note that the choice of $`M_1`$ has no density-related effects on $`N_c`$ and $`L_c`$. ### 5.2. Dynamical Pairs in Redshift Space While it is preferable to identify companions based on their true physical pair separation, this is clearly not feasible when dealing with data from redshift surveys. In the absence of independent distance measurements for each galaxy, one must resort to identifying companions in redshift space. In this section, we outline a straightforward approach for measuring our new pair statistics in redshift space. We then attempt to relate these statistics to their counterparts in real space. For any given pair of galaxies in redshift space, one can compute two basic properties which describe the intrinsic pair separation : the projected physical separation (hereafter $`r_p`$) and the rest-frame relative velocity along the line of sight (hereafter $`\mathrm{\Delta }v`$). For a pair of galaxies with redshifts $`z_i`$ (primary galaxy) and $`z_j`$ (secondary), with angular separation $`\theta `$, these quantities are given by $`r_p=\theta d_A(z_i)`$ and $`\mathrm{\Delta }v=c|z_jz_i|/(1+z_i)`$, where $`d_A(z_i)`$ is the angular diameter distance at redshift $`z_i`$. Note that $`r_p`$ gives the projected separation at the redshift of the primary galaxy. We define a close companion as one in which the separation (both projected and line-of-sight) is less than some appropriate separation, such that $`r_pr_p^{\mathrm{max}}`$ and $`\mathrm{\Delta }v\mathrm{\Delta }v^{\mathrm{max}}`$. The line-of-sight criterion depends on both the physical line-of-sight separation and the line-of-sight peculiar velocity of the companion. It is of course not possible to determine the relative contributions of these components without distance information. However, for the small companion separations we will be concerned with, the peculiar velocity component is likely to be dominant in most cases, as we will be dealing with a field sample of galaxies (the same would not be true in the high velocity environment of rich clusters). Hence, this criterion serves primarily to identify companions with low peculiar velocities. While this is fundamentally different from the pure separation criterion used in real space, it too will serve to identify companions with the highest likelihood of undergoing imminent mergers. Using this definition of a close companion, it is straightforward to compute $`N_c`$ and $`L_c`$, using Equations 6 and 7. Thus, the complexities of redshift space do not greatly complicate the computation of these pair statistics. As in real space, we wish to relate these statistics to measurements on larger scales, given reasonable assumptions about the LF and CF. The situation is more complicated in redshift space, and therefore involves additional assumptions. We stress, however, that these assumptions apply only to the method of relating pair statistics to large scale measures, and not to the measured pair statistics themselves. To outline an algorithm for generating these predictions, we follow the approach of the previous section. We begin by modifying Equations 5.1 and 5.1, integrating over the new pair volume defined in redshift space. In order to do this, we use the two dimensional correlation function in redshift space, $`\xi _2(r_p,r_v,z)`$, giving $`N_{c_i}(z_i,MM_2,r_pr_p^{\mathrm{max}},|r_v|r_v^{\mathrm{max}})=`$ $`n_2(z_i,MM_2){\displaystyle _{r_v^{\mathrm{max}}}^{r_v^{\mathrm{max}}}}{\displaystyle _0^{r_p^{\mathrm{max}}}}[1+\xi _2(r_p,r_v,z_i)]2\pi r_p𝑑r_p𝑑r_v`$ (15) and $`L_{c_i}(z_i,MM_2,r_pr_p^{\mathrm{max}},|r_v|r_v^{\mathrm{max}})=`$ $`j_2(z_i,MM_2){\displaystyle _{r_v^{\mathrm{max}}}^{r_v^{\mathrm{max}}}}{\displaystyle _0^{r_p^{\mathrm{max}}}}[1+\xi _2(r_p,r_v,z_i)]2\pi r_p𝑑r_p𝑑r_v.`$ (16) The two dimensional correlation function is the convolution of the velocity distribution in the redshift direction, $`F(v_z)`$, with the spatial correlation function $`\xi (r,z)`$, given by $$\xi _2(x_p,x_v,z)=_{\mathrm{}}^{\mathrm{}}\xi (\sqrt{x_p^2+y^2},z)F(H(z)[yx_v])𝑑y.$$ (17) Here, $`H(z)`$ is the Hubble constant at redshift $`z`$, given by $`H(z)=H_0(1+z)\sqrt{1+\mathrm{\Omega }_0z}`$. We have ignored the effect of infall velocities, which must be taken into account at larger radii but is an acceptable approximation for small separations. If the form of the CF and LF are known, it is straightforward to integrate Equations 5.2 and 5.2, yielding predictions of $`N_c`$ and $`L_c`$. ### 5.3. Physical Pairs in Redshift Space It is not always possible to have precise redshifts for all galaxies of interest in a sample. A common scenario with redshift surveys is to have redshifts available for a subset of galaxies identified in a flux-limited photometric sample. The photometric sample used to select galaxies for follow-up spectroscopy probes to fainter apparent magnitudes than the spectroscopic sample. In addition, the spectroscopic sample may be incomplete, even at the bright end of the sample. In this section, we will describe the procedure for applying pair statistics to this class of samples. Suppose the primary sample is defined as all galaxies in the spectroscopic sample with absolute magnitudes $`MM_1`$. The secondary sample consists of all galaxies lying in the photometric sample, regardless of whether or not they have measured redshifts. Once again, there may be some overlap between the primary and secondary samples. We must now identify close pairs. For each primary-secondary pair, we can compute $`r_p`$ in precisely the same manner as before (see previous section), since we need only the redshift of the primary galaxy and the angular separation of the pair. However, we are no longer able to compute the relative velocity along the line of sight, since this requires redshifts for both members of the pair. Thus, we do not have enough information to identify close dynamical pairs. However, it is still possible to determine, in a statistical manner, how many physically associated companions are present. This is done by comparing the number (or luminosity) of observed companions with the number (or luminosity) expected in a random distribution. As stressed in Section 5.1, pair statistics depend on the minimum luminosity $`M_2`$ imposed on the secondary sample. While we are now unable to compute the actual luminosity for galaxies in the secondary sample, we must still impose $`M_2`$ if the ensuing pair statistics are to be meaningful. To do this, we make use of the fact that all physical companions must lie at approximately the same redshift as the primary galaxy under consideration. Therefore, $`M_2`$ corresponds to a limiting apparent magnitude $`m_2`$ at redshift $`z_i`$, such that $$m_2=M_2(z_i)5\mathrm{log}h+25+5\mathrm{log}d_L(z_i)+k(z_i),$$ (18) where $`d_L(z_i)`$ is the luminosity distance at redshift $`z_i`$, and $`k(z_i)`$ is the $`k`$-correction. To begin, one finds all observed close companions with $`mm_2(z_i)`$, using only the $`r_p`$ criterion. This results in the quantities $`N_c^\mathrm{D}`$ and $`L_c^\mathrm{D}`$, where the “D” superscript denote companions found in the data sample. One must then estimate the number ($`N_c^\mathrm{R}`$) and luminosity ($`L_c^\mathrm{R}`$) of companions expected at random. The final pair statistics for close physical companions are then given by $`N_c=N_c^\mathrm{D}N_c^\mathrm{R}`$ and $`L_c=L_c^\mathrm{D}L_c^\mathrm{R}`$. We will now describe how to predict these statistics using the known LF and CF. This is relatively straightforward, since the excess $`\xi `$ given by the CF is determined by the relative proportions of real and random companions. The pair statistics are once again integrals over the two dimensional CF in redshift space, as specified by Equations 5.2 and 5.2. In the “$`1+\xi `$” term, the first part gives the random contribution, while the second gives the excess over random. Thus, these pair statistics give the true density of companions, rather than the “excess” density. This is intentional, since mergers will occur even in an uncorrelated, randomly distributed sample of galaxies. At the small separations of interest, usually less than 1% of the correlation length, the difference between the mean density and the mean overdensity is less than about 0.01% in real space. For practical measurements in redshift space, where $`r_v=\mathrm{\Delta }vH(z)^1`$ is of order the correlation length, the background contribution is substantially larger than real space, but still amounts to less than 1%. Thus, for the close pairs considered in this study, it is reasonable to ignore the contribution that random companions make to the sample of physical companions. That is, we take $`1+\xi \xi `$. This allows us to relate the predictions to the measured pair statistics set out above. In principle, Equations 5.2 and 5.2 can be integrated over the range $`\mathrm{}<r_v<\mathrm{}`$ to obtain predictions of $`N_c`$ and $`L_c`$. ### 5.4. Application to Volume-limited Monte Carlo Simulations To illustrate the concepts introduced so far, and to emphasize how these statistics depend on $`M_2`$, we apply these techniques to volume-limited Monte Carlo simulations, which mimic the global distribution of galaxies in the SSRS2 North and South catalogs. Using $`q_0`$=0.5, galaxies are distributed randomly within the co-moving volume enclosed by $`0.005z0.05`$ and the SSRS2 boundaries on the sky (see Section 3). All peculiar velocities are set to zero. To create a volume-limited sample, we impose a minimum luminosity of $`M_B`$=$`16`$, and assign luminosities using the SSRS2 LF (Marzke et al. (1998)), which has Schechter function parameters $`M_{}5\mathrm{log}h`$=$`19.43`$, $`\alpha `$=$`1.12`$, and $`\varphi _{}`$ = $`0.0128h^3\mathrm{Mpc}^3\mathrm{mag}^1`$. An arbitrarily large number of galaxies can be generated, which is of great assistance when looking for small systematic effects. We produce 16000 galaxies in the South, and 8070 in the North; this gives the same density of galaxies in both regions. Using these simulations, we compute $`N_c`$ and $`L_c`$. As these galaxies are distributed randomly (as opposed to real galaxies which are clustered), close pairs are relatively rare. To ensure a reasonable yield of pairs, we use a pair definition of $`r_p^{\mathrm{max}}`$=1 $`h^1`$ Mpc and $`\mathrm{\Delta }v^{\mathrm{max}}`$=1000 km/s. We note that there are no peculiar velocities in these simulations; hence, the $`\mathrm{\Delta }v^{\mathrm{max}}`$ criterion provides upper and lower limits on the line-of-sight distance to companions. Also, recall from the preceding section that the choice of $`M_1`$ has no effect on the pair statistics if clustering is independent of luminosity. Hence, we choose $`M_1`$=$`M_2`$, which maximizes the size of the primary sample, and therefore minimizes the measurement errors in $`N_c`$ and $`L_c`$. With these assumptions, we compute pair statistics for a range of choices of $`M_2`$. Errors are computed using the Jackknife technique. For this resampling method, partial standard deviations, $`\delta _i`$, are computed for each object by taking the difference between the quantity being measuring, $`f`$, and the same quantity with the $`i^{\mathrm{th}}`$ galaxy removed from the sample, $`f_i`$, such that $`\delta _i=ff_i`$. For a sample of $`N`$ galaxies, the variance is given by $`[(N1)/N_i\delta _i^2]^{1/2}`$ (Efron 1981; Efron & Tibshirani 1986). Results are given in Figure 1. Both statistics continue to increase as $`M_2`$ becomes fainter. $`N_c`$ diverges at faint magnitudes, while $`L_c`$ is seen to converge. This behaviour is a direct consequence of the shape of the LF; $`N_c`$ converges for $`\alpha >1`$, while $`L_c`$ converges for $`\alpha >2`$. The existence and magnitude of these trends clearly demonstrate the need to specify $`M_2`$ when measuring pair statistics. ## 6. APPLICATION TO FLUX-LIMITED SAMPLES The preceding section gives a straightforward prescription for computing pair statistics in volume-limited samples. However, redshift surveys are generally flux-limited. By defining a volume-limited sample within such a survey, one must discard a large proportion of the data. In this section, we will outline how these pair statistics can be applied to flux-limited surveys. Pair statistics necessarily depend on both clustering and mean density, as shown by Equations 5.2 and 5.2. In a flux-limited sample, both clustering and mean density will vary throughout the sample. We will use these equations to account for redshift-dependent changes in mean density, and we will demonstrate how to minimize the effects of clustering differences. These techniques will then be tested with Monte Carlo simulations. ### 6.1. Dependence on Clustering By removing the fixed luminosity limit, the overall distribution of galaxy luminosities will vary with redshift within the sample, and the mean luminosity of the sample will differ from the volume-limited sample. However, galaxy clustering is known to be luminosity dependent. Measures of the galaxy correlation function (e.g., Loveday et al. 1995, Willmer et al. 1998), power spectrum (e.g., Vogeley 1993), and counts in cells (Benoist et al. 1996) all find that luminous galaxies ($`L>L_{}`$) are more clustered than sub-$`L_{}`$ galaxies, typically by a factor of $`2`$. This increase in clustering may be particularly strong (factor $`4`$) for very luminous galaxies ($`M_B<21`$). Clearly, this effect should not be ignored when computing pair statistics. In principle, this could be incorporated into the measurement of these pair statistics. However, available pair samples are too small to measure this dependence. We choose instead to minimize these effects by restricting the analysis to a fixed range in absolute magnitude, within which luminosity-dependent clustering is small or negligible. This is done by imposing additional bright ($`M_{\mathrm{bright}}`$) and faint ($`M_{\mathrm{faint}}`$) absolute magnitude limits on the sample. Having thereby reduced the effects of luminosity segregation, we then assume that the remaining differences will not have a significant effect on the measured pair statistics. In Section 8.3, we use the SSRS2 sample to demonstrate empirically that this is in fact a reasonable assumption. ### 6.2. Dependence on Limiting Absolute Magnitude In Section 5, we demonstrated that these pair statistics are meaningful only if one specifies the minimum luminosity of the primary and secondary samples. For a flux-limited sample, however, the minimum luminosity of the sample increases with redshift. One must therefore decide on a representative minimum luminosity, and account for differences between the desired minimum luminosity and the redshift-dependent minimum imposed by the apparent magnitude limit of the sample. If the LF is known, this can be achieved by weighting each galaxy appropriately. In this section, we outline a weighting scheme which makes this correction. #### 6.2.1 Weighting of Secondary Sample Consider a flux-limited sample in which host galaxies are located at a variety of redshifts. Those at low redshift will have the greatest probability of having close companions that lie above the flux limit, since the flux limit corresponds to an intrinsic luminosity that is fainter than that for galaxies at higher redshift. If we wish to avoid an inherent bias in the pair statistics, we must correct for this effect. Furthermore, we must account for any limits in absolute magnitude imposed on the sample to reduce the effects of luminosity-dependent clustering (§ 6.1). Finally, we have demonstrated the importance of specifying a limiting absolute magnitude for companions ($`M_2`$) when computing pair statistics. Therefore, we must attempt to correct the pair statistics to the values that would have been achieved for a volume-limited secondary sample with $`M_{\mathrm{bright}}MM_2`$. Qualitatively, this correction should assign greater importance (or weight) to the more rare companions found at the high redshift end of the flux-limited sample. To make this correction as rigorous as possible, we will use the galaxy LF. By integrating the LF over a given range in absolute magnitude, one can obtain an estimate of the mean number or luminosity density of galaxies in the sample. By performing this integration at any given redshift, accounting for the allowed ranges in absolute magnitude and the flux limit, it is possible to quantify how the mean density varies with redshift within the defined sample. This information can be used to remove this unwanted bias from the pair statistics. We assign a weight to each galaxy in the secondary sample, which renormalizes the sample to the density corresponding to $`M_{\mathrm{bright}}MM_2`$. We must first determine $`M_{\mathrm{lim}}(z_i)`$, which gives the limiting absolute magnitude allowed at redshift $`z_i`$. At most redshifts, this is imposed by the limiting apparent magnitude $`m`$, such that $`M_{\mathrm{lim}}(z_i)=m5\mathrm{log}d_L(z)25k(z)`$. At the low redshift end of the sample, however, $`M_{\mathrm{faint}}`$ (defined in § 6.1) will take over. That is, the limiting absolute magnitude used for identifying galaxies in the secondary sample is given by $$M_{\mathrm{lim}}(z_i)=\mathrm{max}[M_{\mathrm{faint}},m5\mathrm{log}d_L(z)25k(z)].$$ (19) The selection function, denoted $`S(z)`$, is defined as the ratio of densities in flux-limited versus volume-limited samples. This function, given in terms of number density ($`S_N(z)`$) and luminosity density ($`S_L(z)`$), is as follows : $$S_N(z_i)=\frac{_{M_{\mathrm{bright}}}^{M_{\mathrm{lim}}(z_i)}\varphi (M)𝑑M}{_{M_{\mathrm{bright}}}^{M_2}\varphi (M)𝑑M}$$ (20) $$S_L(z_i)=\frac{_{M_{\mathrm{bright}}}^{M_{\mathrm{lim}}(z_i)}\varphi (M)L(M)𝑑M}{_{M_{\mathrm{bright}}}^{M_2}\varphi (M)L(M)𝑑M},$$ (21) where $`L(M)`$ is defined in Equation 13. In order to recover the correct pair statistics, each companion must be assigned weights $`w_{N2}(z_i)=1/S_N(z_i)`$ and $`w_{L2}(z_i)=1/S_L(z_i)`$. The total number and luminosity of close companions for the $`i^{th}`$ primary galaxy, computed by summing over the $`j`$ galaxies satisfying the “close companion” criteria, is given by $`N_{c_i}=_jw_{N2}(z_j)`$ and $`L_{c_i}=_jw_{L2}(z_j)L_j`$ respectively. By applying this weighting scheme to all galaxies in the secondary sample, we will retrieve pair statistics that correspond to a volume-limited sample with $`M_{\mathrm{bright}}MM_2`$. #### 6.2.2 Weighting of Primary Sample The above weighting scheme ensures that the number and luminosity of companions found around each primary galaxy is normalized to $`M_{\mathrm{bright}}MM_2`$. However, these estimates are obviously better for galaxies at the low redshift end of the primary sample, since they will have the largest number of observed companions. Recall that $`N_c`$ and $`L_c`$ are quantities that are averaged over a sample of primary galaxies. In order to minimize the errors in these statistics, we assign weights to the primary galaxies (denoted $`w_{N1}(z_i)`$ and $`w_{L1}(z_i)`$) which are inversely proportional to the square of their uncertainty. If the observed number and luminosity of companions around the $`i^{th}`$ primary galaxy are given by $`N_i(obs)`$ and $`L_i(obs)`$ respectively, and if we assume that the uncertainties are determined by Poisson counting statistics, then $`N_{c_i}=w_{N2}(z_i)N_i(obs)\pm w_{N2}(z_i)\sqrt{N_i(obs)}`$ and $`L_{c_i}=w_{L2}(z_i)L_i(obs)\pm w_{L2}(z_i)\sqrt{L_i(obs)}`$. On average, these quantities will be related to expectation values $`<N_c>`$ and $`<L_c>`$ by $`<N_c>=w_{N2}(z_i)N_i(obs)`$ and $`<L_c>=w_{L2}(z_i)L_i(obs)`$. Combining these relations yields $$w_{N1}(z_i)=\frac{1}{N_i(obs)w_{N2}(z_i)^2}=\frac{1}{<N_c>w_{N2}(z_i)}w_{N2}(z_i)^1$$ (22) $$w_{L1}(z_i)=\frac{1}{L_i(obs)w_{L2}(z_i)^2}=\frac{1}{<L_c>w_{L2}(z_i)}w_{L2}(z_i)^1$$ (23) That is, the optimal weighting is the reciprocal of the weighting scheme used for companions. Therefore, weights $`w_{N1}(z_i)=S_N(z_i)`$ and $`w_{L1}(z_i)=S_L(z_i)`$ should be assigned to primary galaxies. The pair statistics are then computed as follows : $$N_c=\frac{\underset{i}{}w_{N1}(z_i)N_{c_i}}{_iw_{N1}(z_i)}$$ (24) $$L_c=\frac{\underset{i}{}w_{L1}(z_i)L_{c_i}}{_iw_{L1}(z_i)}.$$ (25) It is worth noting that, for a close pair, both galaxies will lie at roughly the same redshift, meaning that $`w_1(z_i)\times w_2(z_j)1`$. We choose not to make this approximation, in order to keep these relations valid for pairs that are not close, and to allow for future application to pairs with additional selection weights. However, we stress that, with or without this approximation, the primary weights in the denominator provide an overall correction for the flux limit, unlike the traditional pair fraction. Note also that, for a volume-limited sample, weights for all galaxies in the primary and secondary samples are equal, reducing these equations to $`N_c=N_{c_i}/N_1`$ and $`L_c=L_{c_i}/N_1`$, as defined in Section 5.1. #### 6.2.3 Boundary Effects A small correction must be made to these weights if a primary galaxy lies close to a region of space that is not covered by the survey. This will happen if a galaxy lies close to the boundaries on the sky, or close to the minimum or maximum redshift allowed. If this is the case, it is possible that close companions will be missed, leading to an underestimate of the pair statistics. Therefore, we must account for these effects. First, we consider galaxies lying close to the survey boundaries on the sky, as defined in Section 3. For each galaxy in the primary sample, we compute the fraction of sky with $`r_p^{\mathrm{min}}r_pr_p^{\mathrm{max}}`$ that lies within the survey boundaries. This fraction will be denoted $`f_b`$. For SSRS2, our usual choices of $`r_p^{\mathrm{min}}`$ and $`r_p^{\mathrm{max}}`$ (see § 7) make this a very small effect, with $`f_b`$=1 for 99.75% of galaxies in the primary sample. Having measured $`f_b`$ for each galaxy in the primary sample, we must incorporate this into the measurement of the pair statistics. The first task is to ensure that we correct the number of companions to match what would be expected if coverage was complete. We do this by assigning each companion a boundary weight $`w_{b_2}`$ = $`1/f_b`$, where $`f_b`$ is associated with its host galaxy from the primary sample. By multiplying each companion by its boundary weight, we will recover the correct number of companions. We must also adjust weights for the primary galaxies. Following the method described in the previous section, we wish to give less weight to galaxies that are likely to have fewer observed companions. Therefore, each primary galaxy is assigned a boundary weight $`w_{b_1}`$ = $`f_b`$. We now consider galaxies which lie near the survey boundaries along the line of sight. If a primary galaxy lies close to the minimum or maximum redshift allowed, it is possible that we will miss companions because they lie just across this redshift boundary. In order to account correctly for this effect, one would need to model the velocity distribution of companions. As this requires several assumptions, we choose instead to exclude all companions that lie between a primary galaxy and its nearest redshift boundary, provided the boundary lies within $`\mathrm{\Delta }v^{\mathrm{max}}`$ of the primary galaxy. To account for this exclusion, we assume that the velocity distribution is symmetric along the line of sight. Thus, as we will miss half of the companions for these galaxies, we assign a weight of $`w_{2_v}`$=2 to any companions found in the direction opposite to the boundary. We must also consider how to weight the primary galaxies themselves. Clearly, primary galaxies close to the redshift boundaries will be expected to have half as many observed companions as other primary galaxies. To minimize the errors in computing the pair statistics, we assign these primary galaxies weights $`w_{v_1}`$=0.5. To summarize, weights for companions in the secondary sample are given by $$w_{N_2}=S_N(z_i)^1w_{b_2}w_{v_2},$$ (26) $$w_{L_2}=S_L(z_i)^1w_{b_2}w_{v_2},$$ (27) while primary galaxies are assigned weights $$w_{N_1}=S_N(z_i)w_{b_1}w_{v_1},$$ (28) $$w_{L_1}=S_L(z_i)w_{b_1}w_{v_1}.$$ (29) ### 6.3. Confirmation Using Monte Carlo Simulations We will now perform a test to see if this weighting scheme achieves the desired effects. To do this, we will use flux-limited Monte Carlo simulations, for which the intrinsic density and clustering are fixed. Therefore, the intrinsic pair statistics do not depend on redshift or luminosity. If the secondary sample weights are correct, the measured pair statistics will be the same everywhere (within the measurement errors), regardless of redshift or luminosity. We will also check to see if the weights for the primary sample are correct. If they are, the errors on the pair statistics will be minimized, as desired. The flux-limited Monte Carlo simulations were generated in a similar manner to the simulations described in Section 5.4; however, a limiting apparent magnitude of $`m_B15.5`$ was imposed. Sample sizes of 8000 (South) and 4035 (North) were used, providing a good match to the overall density in SSRS2. The resulting simulations are similar to SSRS2 in all respects, except for the absence of clustering. We have already established how the pair statistics depend on the choice of $`M_1`$ and $`M_2`$. In the following analysis, we choose $`M_2`$=$`M_1`$=$`19`$. In Section 6.1, we outlined reasons for restricting the sample to a fixed range in absolute magnitude. Here, we demonstrate how the chosen range affects $`N_c`$ and $`L_c`$. For comparison, we also compute $`N_c`$ without normalizing to a specified range in absolute magnitude (in this case, $`w_{N_1}`$=$`w_{N_1}`$=1). This provides some insight into the behaviour of the traditional (uncorrected) pair fraction. These tests are most straightforward if the intrinsic pair statistics are the same everywhere in the enclosed volume. This is not quite true for these simulations, however. Galaxies are distributed randomly within the enclosed co-moving volume. As a result, the physical density varies with redshift as $`(1+z)^3`$. In addition, the volume element encompassed by the line-of-sight pair criterion $`\mathrm{\Delta }v`$ varies with redshift as $`(1+z)^{3/2}`$ for $`q_0`$=0.5. In order to have the simulations mimic a sample with universal pair statistics, we normalize the sample for these effects by weighting each galaxy by $`(1+z)^{3/2}`$. We stress that this is done only for the Monte Carlo simulations. One should not apply either of these corrections to real redshift data. In Figure 2, the pair statistics are computed for a range of $`M_{\mathrm{faint}}`$. In addition, we compute $`N_c`$ without weighting by the luminosity function, to demonstrate the danger of ignoring this important correction. This statistic is directly analagous to the traditional (uncorrected) pair fraction used in the literature. It is clear that both $`N_c`$ and $`L_c`$ are independent of the choice of $`M_{\mathrm{faint}}`$, within the errors. This verifies that we have correctly accounted for the biases introduced by the apparent magnitude limit. In contrast, the unweighted $`N_c`$ is seen to have a strong dependence on $`M_{\mathrm{faint}}`$. As expected, it increases as $`M_{\mathrm{faint}}`$ becomes fainter, due to the increase in sample density. We stress that this does not happen with the normalized $`N_c`$ and $`L_c`$ statistics, because both are corrected to a fixed range in limiting absolute magnitude. Finally, we demonstrate that the weighting scheme used for the primary sample (§ 6.2.2) does in fact minimize errors in $`N_c`$ and $`L_c`$. Recall that the weighting used was the reciprocal of the weights for the secondary sample. Here we will assume that $`w_{N1}w_{N2}(z_i)^x`$ and $`w_{L1}w_{L2}(z_i)^x`$. In Section 6.2.2, justification was given for setting $`x`$=$`1`$. Here, we will allow $`x`$ to vary, in order to investigate empirically which value minimizes the errors. Special cases of interest are $`x`$=0 (no weighting) and $`x`$=1 (same weighting as secondary sample). The results are given in Figure 3. The relative errors in $`N_c`$ and $`L_c`$ reach a minimum at $`x1`$, as expected. Errors are $``$ 40% larger if no weighting is used ($`x`$=0). For $`x`$=1, errors are much larger, increasing by nearly a factor of 5. While errors increase dramatically for $`x0`$, they change slowly around $`x`$=$`1`$. Clearly, $`x`$ =$`1`$ is an excellent choice. ## 7. A PAIR CLASSIFICATION EXPERIMENT The first step in applying these techniques to a real survey of galaxies is to decide on a useful close pair definition. This involves imposing a maximum projected physical separation ($`r_p^{\mathrm{max}}`$) and, if possible, a maximum line-of-sight rest-frame velocity difference ($`\mathrm{\Delta }v^{\mathrm{max}}`$). The limits should be chosen so as to extract information on mergers in an optimal manner. This involves a compromise between the number and merging likelihood of pairs. While one should focus on companions which are most likely to be involved in mergers, a very stringent pair definition may yield a small and statistically insignificant sample. In previous close pair studies, the convention has been to set $`r_p^{\mathrm{max}}`$ = 20 $`h^1`$ kpc. Pairs with separations of $`r_p`$ 20 $`h^1`$ kpc are expected to merge within 0.5 Gyr (e.g., Barnes (1988), Patton et al. (1997)). We note, however, that timescale estimates are approximate in nature, and have yet to be verified. In earlier work, it has not been possible to apply a velocity criterion, since redshift samples have been too small to yield useful pair statistics using only galaxies with measured redshifts. Instead, all physical companions have been used, with statistical correction for optical contamination (Patton et al. (1997)). With a complete redshift sample, we can improve on this. This can be seen by inspecting a plot of $`r_p`$ versus $`\mathrm{\Delta }v`$ for the SSRS2 pairs, given in Figure 4. By imposing a velocity criterion, we can eliminate optical contamination; furthermore, we are able to concentrate on the physical pairs with the lowest relative velocities, and hence the greatest likelihood of merging. We can now use our large sample of low-$`z`$ pairs to shed new light on these issues. We will use images of these pairs in an attempt to determine how signs of interactions are related to pair separation. We begin by finding all 255 SSRS2 pairs with $`r_p100`$ $`h^1`$ kpc, computing $`r_p`$ and $`\mathrm{\Delta }v`$ for each. Images for these pairs were extracted from the Digitized Sky Survey. Interactions were immediately apparent in some of these pairs, and the images were deemed to be of sufficient quality that a visual classification scheme would be useful. An interaction classification parameter (I<sub>c</sub>) was devised, where I<sub>c</sub>=0 indicates that a given pair is “definitely not interacting”, and I<sub>c</sub>=10 indicates “definitely interacting”. In order to avoid a built-in bias, the classifier is not given the computed values of $`r_p`$ and $`\mathrm{\Delta }v`$. The classifier uses all visible information available (tidal tails and bridges, distortions/asymmetries in member galaxies, apparent proximity, etc.). Classifications were performed by three of us (DRP, ROM, RGC), and the median classification was determined for each system. The results are presented in Figure 5. A clickable version of this plot, which allows the user to see the corresponding Digitized Sky Survey image for each pair, is available at http://www.astro.utoronto.ca/$``$patton/ssrs2/Ic. There are several important features in this plot. First, there is a clear correlation between $`I_c`$ and $`r_p`$, with closer pairs exhibiting stronger signs of interactions. There are several interacting pairs with $`r_p50`$ $`h^1`$ kpc. While these separations are fairly large, it is not surprising that there would be some early-stage mergers with these separations (e.g., Barton, Bromley, & Geller 1998). An excellent example of this phenomenon is the striking tail-bridge system Arp 295a/b (cf. Hibbard & van Gorkom (1996)), which has $`r_p`$ = 95 $`h^1`$ kpc. However, these systems clearly do not dominate; almost all pairs with large separations have very low interaction classifications. The majority of pairs showing clear signs of interactions/mergers have $`r_p<20`$ $`h^1`$ kpc . There is also a clear connection with $`\mathrm{\Delta }v`$. Pairs with $`\mathrm{\Delta }v>600`$ km/s do not exhibit signs of interactions, with 61/63 (97%) classified as $`I_c1`$. This indicates that interactions are most likely to be seen in low velocity pairs, as expected. We note, however, that there are very few optical pairs (i.e., small $`r_p`$ and large $`\mathrm{\Delta }v`$) in this low redshift sample. At higher redshift, increased optical contamination may lead to difficulties in identifying interacting systems when the galaxies are close enough to have overlapping isophotes. Clearly, it is necessary to have redshift information for both members of each pair if one is to exclude these close optical pairs. After close inspection of Figure 5, we decided on close pair criteria of $`r_p^{\mathrm{max}}`$ = 20 $`h^1`$ kpc and $`\mathrm{\Delta }v^{\mathrm{max}}`$ = 500 km/s. A mosaic of some of these pairs is given in Figure 6. In this regime, 31% (9/29) exhibit convincing evidence for interactions (I<sub>c</sub> $`9`$), while 69% (20/29) show some indication of interactions (I<sub>c</sub> $`6`$). Furthermore, the vast majority (9/10) of pairs with clear signs of interactions (I<sub>c</sub> $`9`$) are found in this regime. These criteria appear to identify a sample of pairs which are likely to be undergoing mergers; moreover, the resulting sample includes most of the systems classified as interacting. We also impose an inner boundary of $`r_p`$ = 5 $`h^1`$ kpc. This limit is chosen so as to avoid the confusion that is often present on the smallest scales. In this regime, it is often difficult to distinguish between small galaxies and sub-galactic units, particularly in merging systems. While we are omitting the most likely merger candidates, those at separations $`<`$ 5 $`h^1`$ kpc are not expected to account for more than $``$ 5% of the companions within 20 $`h^1`$ kpc . This expectation, which has yet to be verified, is based both on pair counts in HST imaging (Burkey et al. (1994)) and on inward extrapolation of the correlation function (Patton et al. (1997)). While this inner boundary will lead to a slight decrease in $`N_c`$ and $`L_c`$, it should have no significant effect on estimates of merger/accretion rate evolution, provided the same restriction is applied to comparison samples at other redshifts. ## 8. APPLICATION TO SSRS2 In the preceding sections, we have outlined techniques for measuring pair statistics in a wide variety of samples. We have demonstrated a robust method of applying this approach to flux-limited samples, accounting for redshift-dependent density changes and minimizing differences in clustering. We have also selected pair definitions that identify the most probable imminent mergers. We will now apply these techniques to the SSRS2 survey. As this is a complete redshift survey, redshifts are available for all close companions; hence, for the first time, we will measure pair statistics using only close dynamical pairs. After limiting the analysis to a reasonable range in absolute magnitude, we compute $`N_c`$ and $`L_c`$ for the SSRS2 survey. ### 8.1. Defining Survey Parameters In Section 6.1, we emphasized the importance of restricting the sample in absolute magnitude, to minimize bias due to luminosity-dependent clustering. For SSRS2, we first impose a bright limit of $`M_{\mathrm{bright}}`$ = $`21`$. All galaxies brighter than this are hereafter excluded from the analysis. This allows us to avoid the most luminous galaxies, which are probably the most susceptible to luminosity-dependent clustering; however, this reduces the size of the sample by only 0.5%. We also impose a faint absolute magnitude limit of $`M_{\mathrm{faint}}`$ = $`17`$, which results in the exclusion of intrinsically faint galaxies at $`z<0.01`$. This guards against the possibility that these intrinsically faint galaxies are clustered differently than the bulk of the galaxies in the sample. This pruning of the sample is illustrated in Figure 7. These restrictions allow us to minimize concerns about luminosity-dependent clustering while retaining 90% of the sample. The final results are insensitive to these particular choices (see Section 8.3). The above limits in absolute magnitude, along with the flux limit, define the usable sample of galaxies. In order to compute pair statistics, we must also normalize the measurements to a given range in absolute magnitude, for both the primary and secondary samples. The mean limiting absolute magnitude of the primary sample, weighted according to Section 6.2.2, is $`M_B`$ = $`17.8`$. For convenience, we set $`M_2(B)`$ = $`18.0`$ (we will compute pair statistics for $`19M_2(B)17`$ in the following section). For reference, we note that this corresponds to $`m_B`$=$`15.5`$ at $`z`$=0.017. As we are dealing with a complete redshift sample, we set $`M_1(B)`$=$`M_2(B)`$ in order to use all of the available information. Finally, as we have limited the sample using $`M_{\mathrm{bright}}`$ = $`21`$, this will be used in conjunction with $`M_2(B)`$ to derive pair statistics for galaxies with $`21M_B18`$. ### 8.2. SSRS2 Pair Statistics Using these parameters, we identified all close companions in SSRS2. The North sample yielded 27 companions, and 53 were found in the South, giving a total of 80. We emphasize that it is companions that are counted, rather than pairs; hence, if both members of a pair fall within the primary sample, the pair will usually yield 2 companions. A histogram of companion absolute magnitudes is given in Figure 8. This plot shows that 90% of the companions we observe in our flux-limited sample fall in the range $`21M_B18`$. Hence, galaxies with $`18M_B17`$ do not dominate the sample. Tables 2 and 3 give complete lists of close aggregates (pairs and triples) for SSRS2 North and South respectively. These systems contain all companions used in the computation of pair statistics. These tables list system ID, number of members, $`r_p`$ ($`h^1`$ kpc), $`\mathrm{\Delta }v`$ (km/s), RA (1950.0), DEC (1950.0), and recession velocity (km/s). DSS images for these systems were given earlier in Figure 6. Using this sample of companions, the pair statistics were computed. The results are given in Table 1. Errors were computed using the Jackknife technique. Results from the two subsamples were combined, weighting by Jackknife errors, to give $`N_c=0.0226\pm 0.0052`$ and $`L_c=0.0216\pm 0.0055\times 10^{10}h^2L_{}`$ at $`z`$ = 0.015. Results from the two subsamples agree within the quoted 1$`\sigma `$ errors. To facilitate future comparison with other samples, we also generate pair statistics spanning the range $`19M_2(B)17`$ (see Table 4). We note, however, that while we account for changes in number and luminosity density over this luminosity range (using LF weights described in Section 6), there is no correction for changes in clustering. Hence, our statistics should be considered most appropriate for $`M_2(B)`$ = $`18`$, and more approximate in nature at brighter and fainter levels. The results in Table 4 indicate that $`N_c`$ increases by a factor of 5 between $`M_2(B)`$ = $`19`$ and $`M_2(B)`$ = $`17`$, resulting solely from an increase in mean number density. The change in $`L_c`$ is less pronounced, with an increase by a factor of 2 over the same luminosity range. These substantial changes in both statistics emphasize the need to specify $`M_2`$ when computing pair statistics and comparing results from different samples. In addition, the smaller change in $`L_c`$ is indicative of the benefits of using a luminosity statistic such as $`L_c`$, which is more likely to converge as one goes to fainter luminosities (see Section 5.4). $`L_c`$ will always converge faster than $`N_c`$, thereby reducing the sensitivity to $`M_2`$. Furthermore, it is possible to retrieve most of the relevant luminosity information without probing to extremely faint levels. For example, for the SSRS2 LF, 70% of the total integrated luminosity density is sampled by probing down to $`M_2(B)`$ = $`18`$. To first order, the same will be true for $`L_c`$. Going 2 magnitudes fainter would increase the completeness to 95%. While we are currently unable to apply pair statistics down to these faint limits, this will be pursued when deeper surveys become available. ### 8.3. Sensitivity of Results to Survey Parameters In this section, we explore the effects of choosing different survey parameters. Earlier in this study, we demonstrated that $`N_c`$ and $`L_c`$ are insensitive to the choice of survey limits in absolute magnitude, provided clustering is independent of luminosity and the pair statistics are normalized correctly. Here, we test this hypothesis empirically. First, we compute the pair statistics for a range in $`M_{\mathrm{faint}}`$, normalizing the statistics to $`21M_B18`$ in each case. Figure 9 demonstrates a possible trend of decreasing pair statistics with fainter $`M_{\mathrm{faint}}`$. This trend, however, is significant only for the brightest galaxies ($`M_{\mathrm{faint}}<19`$). This is consistent with the findings of Willmer et al. (1998), who measure an increase in clustering for bright galaxies in SSRS2, on scales of $`r_p>1`$ $`h^1`$ Mpc. For fainter $`M_{\mathrm{faint}}`$, there is no significant dependence. The pair statistics vary by $``$ 5% over the range $`17.5M_{\mathrm{faint}}16.5`$, which is well within the error bars. Therefore, we conclude that our choice of $`M_{\mathrm{faint}}`$ = $`17`$ has a negligible effect on $`N_c`$ and $`L_c`$. This implies that, to first order, clustering is independent of luminosity within this sample. Next, we investigate how the pair statistics depend on our particular choices of $`r_p^{\mathrm{max}}`$ and $`\mathrm{\Delta }v^{\mathrm{max}}`$, which comprise our definition of a close companion. First, we compute pair statistics for 10 $`h^1`$ kpc$`r_p^{\mathrm{max}}100`$$`h^1`$ kpc, with $`\mathrm{\Delta }v^{\mathrm{max}}=500`$ km/s. Results are given in Figure 10. This plot indicates a smooth increase in both statistics with $`r_p^{\mathrm{max}}`$. This trend is expected from measurements of the galaxy CF. The CF is commonly expressed as a power law of the form $`\xi (r,z)=(r_0/r)^\gamma `$, with $`\gamma `$=1.8 (Davis & Peebles (1983)). Integration over this function yields pair statistics that vary as $`r_p^{3\gamma }r_p^{1.2}`$, which is in good agreement with the trend found in Figure 10. From this plot, it also appears likely that there are systematic differences between the two subsamples. This is hardly surprising, since there are known differences in density between the subsamples, and it is likely that there are non-negligible differences in clustering as well. This cosmic variance is not currently measurable on the smaller scales ($`r_p20`$ $`h^1`$ kpc) relevant to our main pair statistics. Hence, we choose to ignore these differences for now. However, these field-to-field variations are certain to add some systematic error to our quoted pair statistics. We also compute pair statistics for a range in $`\mathrm{\Delta }v^{\mathrm{max}}`$. This is done first for $`r_p^{\mathrm{max}}=20`$ $`h^1`$ kpc, showing the relative contributions at different velocities to the main pair statistics quoted in this paper. We also compute statistics using $`r_p^{\mathrm{max}}=100`$$`h^1`$ kpc, in order to improve the statistics. Results are given in Figure 11. Several important conclusions may be drawn from this plot. First, at small velocities ($`\mathrm{\Delta }v^{\mathrm{max}}<700`$ km/s), both pair statistics increase with $`\mathrm{\Delta }v^{\mathrm{max}}`$, as expected. This simply indicates that one continues to find additional companions as the velocity threshold increases. Secondly, it appears that our choice of $`\mathrm{\Delta }v^{\mathrm{max}}`$ was a good one. The $`r_p20`$ $`h^1`$ kpc pair statistics increase very little beyond $`\mathrm{\Delta }v^{\mathrm{max}}=500`$ km/s, while the contamination due to non$``$merging pairs would continue to increase (see Figure 4). Moreover, as both pair statistics flatten out at around $`\mathrm{\Delta }v^{\mathrm{max}}=500`$ km/s, small differences in the velocity distributions of different samples should not result in large differences in their pair statistics. Finally, for $`r_p^{\mathrm{max}}100`$ $`h^1`$ kpc, the pair statistics continue to increase out to $`2000`$ km/s. This indicates an increase in velocity dispersion at these larger separations. This provides additional confirmation that one is less likely to find low$``$velocity pairs at larger separations, thereby implying that mergers should also be less probable. ## 9. DISCUSSION ### 9.1. Comparison with Earlier Estimates of the Local Pair Fraction All published estimates of the local pair fraction have been hindered by small sample sizes and a lack of redshifts. In addition, as demonstrated throughout this paper, the traditional pair fraction is not a robust statistic, particularly when applied to flux-limited surveys. The new statistics introduced in this paper, along with careful accounting for selection effects such as the flux limit, yield the first secure measures of pair statistics at low redshift. Therefore, strictly speaking, the results in this paper cannot be compared directly with earlier pair statistics. However, it is possible to check for general consistency in results, and we will attempt to do so. As discussed in Section 2, Patton et al. (1997) estimated the local pair fraction to be $`4.3\pm 0.4\%`$, using the UGC catalog. The Patton et al. (1997) estimate was based on a flux-limited sample with $`B14.5`$, and a mean redshift of $`z`$=0.0076. This corresponds roughly to an average limiting absolute magnitude of $`M_B`$ = $`17.3`$. Loosely speaking, this is analogous to $`M_2`$. The pair definition used in their estimate was $`r_p20`$ $`h^1`$ kpc, with no $`\mathrm{\Delta }v`$ criterion. $`N_c`$ may be interpreted as an approximation to the traditional pair fraction, provided the relative proportion of triples is small. We recompute the SSRS2 pair statistics, using $`\mathrm{\Delta }v^{\mathrm{max}}=1000`$ km/s in an attempt to match the results that would be found using no $`\mathrm{\Delta }v`$ criterion (see Figure 11). We find $`N_c=0.026\pm 0.006`$. This implies a local pair fraction of $`2.6\pm 0.6\%`$. This value is somewhat smaller than the earlier result, with larger errors. We strongly emphasize that, while these results are broadly consistent, we would not expect excellent agreement, due to the improved techniques used in this study. ### 9.2. The Merger Fraction at $`z0`$ The quantities $`N_c`$ and $`L_c`$ are practical measures of the average numbers and luminosities of companions with relatively high merging probabilities. However, the ambiguity of redshift space is such that some of these companions can be entirely safe from ever merging. That is, $`\mathrm{\Delta }v=500`$ km/s can correspond either to two galaxies at a small physical separation with a large infall velocity, or, to two galaxies at a separation of 5$`h^1`$ Mpc, with no relative peculiar velocity. In order to transform from $`N_C`$ (and $`L_C`$) to an estimate of the incidence of mergers, we must determine what fraction of our close companions have true 3-dimensional physical separations of $`r<20`$ $`h^1`$ kpc. This quantity, which we refer to as $`f_{3\mathrm{D}}`$, has been discussed previously by Yee and Ellingson (1995). This quantity needs to be evaluated based on the small separation clustering and kinematics of the galaxy population, $`r_0`$, $`\gamma `$, and $`\sigma _{12}`$, the parameters $`\mathrm{\Delta }v`$ and $`r_p^{\mathrm{max}}`$, and the projected separation at which two galaxies “optically overlap” in the image. The quantity $`f_{3\mathrm{D}}`$ is then evaluated with a triple integral, first placing the correlation function into redshift space, then integrating over projected and velocity separation. For reasonable choices of pair selection parameters, the outcome is fairly stable at $`f_{3\mathrm{D}}0.5`$. The most important parameter is $`\mathrm{\Delta }V/\sigma _{12}`$, which should lie in the range of 2 to 4. If this parameter is too small, physical pairs will be missed; if it is too large, too many distant companions will be incorporated. Other reasonable choices include setting the ratio of the overlap separation to maximum pair separation to be at least three, and the ratio of the maximum pair separation to correlation length to be at least a factor of 30. We will take $`f_{3\mathrm{D}}`$=0.5 to be the best estimate currently available. We can now estimate the merger fraction ($`f_{\mathrm{mg}}`$) at the present epoch. In this study, we have found $`N_C`$=0.0226$`\pm 0.0052`$. As most companions are found in pairs, rather than triplets or higher order N-tuples, this is comparable to the fraction of galaxies in close pairs. From our estimate of $`f_{3\mathrm{D}}`$, we infer that half of these galaxies are in merging systems, yielding $`f_{\mathrm{mg}}0.011`$. This implies that approximately 1.1% of $`21M_B18`$ galaxies are undergoing mergers at the present epoch. We stress here that this result applies only to galaxies within the specified absolute magnitude limits. Probing to fainter luminosities would cause $`f_{\mathrm{mg}}`$ to increase substantially. In addition, this result applies only to the close companions defined in this analysis. Clearly, by modifying this definition (and therefore changing the typical merger timescale under consideration), the merger fraction would also be certain to change. ### 9.3. The Merger Timescale We now have an idea of how prevalent ongoing mergers are at the present epoch. In order to relate this result to the overall importance of mergers, we must estimate the merger timescale ($`T_{\mathrm{mg}}`$). We will use the properties of our SSRS2 pairs to estimate the mean dynamical friction timescale for pairs in our sample. Following Binney and Tremaine (1987), we assume circular orbits and a dark matter density profile given by $`\rho (r)r^2`$. The dynamical friction timescale $`T_{\mathrm{fric}}`$ (in Gyr) is given by $$T_{\mathrm{fric}}=\frac{2.64\times 10^5r^2v_c}{M\mathrm{ln}\mathrm{\Lambda }},$$ (30) where $`r`$ is the initial physical pair separation in kpc, $`v_c`$ is the circular velocity in km/s, $`M`$ is the mass ($`M_{}`$), and $`\mathrm{ln}\mathrm{\Lambda }`$ is the Coulomb logarithm. We estimate $`r`$ and $`v_c`$ using the pairs in Tables 2 and 3. The mean projected separation is $`r_p14`$$`h^1`$ kpc. As our procedure already includes a correction from projected separation ($`r_p`$) to 3-dimensional separation ($`r`$), we take $`r=r_p`$. Assuming $`h`$=0.7, this leads to $`\overline{r}20`$ kpc. The mean line of sight velocity difference is $`\mathrm{\Delta }v150`$ km/s. We assume the velocity distribution is isotropic, which implies that $`v_c=\sqrt{3}\mathrm{\Delta }v260`$ km/s. The mean absolute magnitude of companions is $`M_B19`$ (see Figure 8). We assume a representative estimate of the galaxy mass-to-light ratio of $`M/L5`$, yielding a mean mass of $`M3\times 10^{10}M_{}`$. Finally, Dubinski, Mihos, and Hernquist (1999) estimate $`\mathrm{ln}\mathrm{\Lambda }2`$, which fits the orbital decay of equal mass mergers seen in simulations. Using Equation 30, we find $`T_{\mathrm{fric}}0.5`$ Gyr. We caution that this estimate is an approximation, and is averaged over systems with a wide range in merger timescales. Nevertheless, we will take $`T_{\mathrm{mg}}`$ = 0.5 Gyr as being representative of the merger timescale for the pairs in our sample. ### 9.4. The Cumulative Effect of Mergers Since $`z1`$ Now that we have estimated the present epoch merger fraction and the merger timescale, we will attempt to ascertain what fraction of present galaxies have undergone mergers in the past. These galaxies can be classified as merger remnants; hence, we will refer to this fraction as the remnant fraction ($`f_{\mathrm{rem}}`$). We begin by imagining the state of affairs at a lookback time of $`t=T_{\mathrm{mg}}`$. Suppose the merger fraction at the corresponding redshift is given by $`f_{\mathrm{mg}}(z)`$. In the time interval between then and the present, a fraction $`f_{\mathrm{mg}}(z)`$ of galaxies will undergo mergers, yielding $`0.5f_{\mathrm{mg}}(z)`$ merger remnants. Therefore, the remnant fraction at the present epoch is given by $$f_{\mathrm{rem}}=1\frac{1f_{\mathrm{mg}}(z)}{10.5f_{\mathrm{mg}}(z)}.$$ (31) Similarly, if we extend this to a lookback time of $`NT_{\mathrm{mg}}`$, where $`N`$ is an integer, then the remnant fraction is given by $$f_{\mathrm{rem}}=1\underset{j=1}{\overset{N}{}}\frac{1f_{\mathrm{mg}}(z_j)}{10.5f_{\mathrm{mg}}(z_j)},$$ (32) where $`z_j`$ corresponds to a lookback time of $`t=jT_{\mathrm{mg}}`$. We now make the simple assumption that the merger rate does not change with time. In this case, our present epoch estimate of the merger fraction holds at all redshifts, giving $`f_{\mathrm{mg}}(z)`$=0.011. In order to convert between redshift and lookback time, we must specify a cosmological model. We assume a Hubble constant of $`h`$=0.7. For simplicity, we assume $`q_0`$=0.5. Therefore, $`z=(13H_0t/2)^{2/3}1`$. Using our merger timescale estimate of $`T_{\mathrm{mg}}`$=0.5 Gyr, we can now investigate the cumulative effect of mergers. With the chosen cosmology, $`z`$=1 corresponds to a lookback time of $``$ 6 Gyr, or 12$`T_{\mathrm{mg}}`$ ($`N`$=12). With this lookback time, Equation 32 yields $`f_{\mathrm{rem}}`$=0.066. This implies that $``$ 6.6% of galaxies with $`21M_B18`$ have undergone mergers since $`z1`$. If the mergers taking place in our sample produce elliptical galaxies, it is worthwhile comparing the remnant fraction to the elliptical fraction (cf. Toomre 1977). The elliptical fraction for bright field galaxies is generally found to be about 10% (e.g., Dressler 1980, Postman and Geller 1984). This result is broadly consistent with the remnant fraction found in this study. While our estimate of the remnant fraction is based on our statistically secure measurement of $`N_C`$, it also relies on fairly crude assumptions regarding the merger fraction and merger timescale. In particular, the merger rate has been assumed to be constant. There is no physical basis for this assumption; in fact, a number of studies have predicted a rise in the merger rate with redshift. If this is true, we will have underestimated the remnant fraction, and the relative importance of mergers. In a future paper (Patton et al. 2000), we will address this issue by investigating how the merger rate changes with redshift. ## 10. CONCLUSIONS We have introduced two new pair statistics, $`N_c`$ and $`L_c`$, which are shown to be related to the galaxy merger and accretion rates respectively. Using Monte Carlo simulations, these statistics are found to be robust to the redshift-dependent density changes inherent in flux-limited samples; this represents a very significant improvement over all previous estimators. In addition, we provide a clear prescription for relating $`N_c`$ and $`L_c`$ to the galaxy CF and LF, enabling straightforward comparison with measurements on larger scales. These statistics are applied to the SSRS2 survey, providing the first statistically sound measurements of pair statistics at low redshift. For an effective range in absolute magnitude of $`21M_B18`$, we find $`N_c=0.0226\pm 0.0052`$ at $`z`$=0.015, implying that $``$ 2.3% of these galaxies have companions within a projected physical separation of 5 $`h^1`$ kpc $`r_p`$ 20 $`h^1`$ kpc and 500 km/s along the line of sight. If this pair statistic remains fixed with redshift, simple assumptions imply that $``$ 6.6% of present day galaxies with $`21M_B18`$ have undergone mergers since $`z`$=1. For our luminosity statistic, we find $`L_c=0.0216\pm 0.0055\times 10^{10}h^2L_{}`$. This statistic gives the mean luminosity in companions, per galaxy. Both of these statistics will serve as local benchmarks in ongoing and future studies aimed at detecting redshift evolution in the galaxy merger and accretion rates. It is our hope that these techniques will be applied to a wide range of future redshift surveys. As we have demonstrated, one must carefully account for differences in sampling effects between pairs and field galaxies. This will be of increased importance when applying pair statistics at higher redshift, as $`k`$-corrections, luminosity evolution, band-shifting effects, and spectroscopic completeness have to be properly accounted for. The general approach outlined in this paper indicates the steps that must be taken to allow for a fair comparison between disparate surveys at low and high redshift. These techniques are applicable to redshift surveys with varying degrees of completeness, and are also adaptable to redshift surveys with additional photometric information, such as photometric redshifts, or even simply photometric identifications. Finally, this approach can be used for detailed studies of both major and minor mergers. We wish to thank all members of the SSRS2 colloboration for their work in compiling the SSRS2 survey, and for making these data available in a timely manner. Digitized Sky Survey images used in this research were obtained from the Canadian Astronomical Data Centre (CADC), and are based on photographic data of the National Geographic Society Palomar Observatory Sky Survey (NGS-POSS). The Digitized Sky Surveys were produced at the Space Telescope Science Institute under U.S. Government grant NAG W-2166. This work was supported by the Natural Sciences and Engineering Research Council of Canada, through research grants to R.G.C. and C.J.P..
warning/0001/gr-qc0001066.html
ar5iv
text
# Reconstruction of a scalar-tensor theory of gravity in an accelerating universe \[ ## Abstract The present acceleration of the Universe strongly indicated by recent observational data can be modeled in the scope of a scalar-tensor theory of gravity. We show that it is possible to determine the structure of this theory (the scalar field potential and the functional form of the scalar-gravity coupling) along with the present density of dustlike matter from the following two observable cosmological functions: the luminosity distance and the linear density perturbation in the dustlike matter component as functions of redshift. Explicit results are presented in the first order in the small inverse Brans-Dicke parameter $`\omega ^1`$. preprint: CPT-99/P.3917 \] Recent observational data on type Ia supernovae explosions at high redshifts $`z\frac{a(t_0)}{a(t)}11`$ obtained independently by two groups , as well as numerous previous arguments (see the recent reviews ), strongly support the existence of a new kind of matter in the Universe whose energy density not only is positive but also dominates the energy densities of all previously known forms of matter \[here $`a(t)`$ is the scale factor of the Friedmann-Robertson-Walker (FRW) isotropic cosmological model, and $`t_0`$ is the present time\]. This form of matter has a strongly negative pressure and remains unclustered at all scales where gravitational clustering of baryons and (non-baryonic) cold dark matter (CDM) is seen. Its gravity results in the present acceleration of the expansion of the Universe: $`\ddot{a}(t_0)>0`$. In a first approximation, this kind of matter may be described by a constant $`\mathrm{\Lambda }`$-term in the gravity equations as first introduced by Einstein. However, a $`\mathrm{\Lambda }`$-term could also be slowly varying with time. If so, this will be soon determined from observational data. In particular, if we use the simplest model of a variable $`\mathrm{\Lambda }`$-term (also called quintessence in ) borrowed from the inflationary scenario of the early Universe, namely an effective scalar field $`\mathrm{\Phi }`$ with some self-interaction potential $`U(\mathrm{\Phi })`$ minimally coupled to gravity, then the functional form of $`U(\mathrm{\Phi })`$ can be determined from observational cosmological functions: either from the luminosity distance $`D_L(z)`$ , or from the linear density perturbation in the dustlike component of matter in the Universe $`\delta _m(z)`$ for a fixed comoving smoothing radius . However, this model cannot account for any future observational data, in particular, for any functional form of $`D_L(z)`$. This happens because a variable $`\mathrm{\Lambda }`$-term in this model should satisfy the weak-energy condition $`\epsilon _\mathrm{\Lambda }+p_\mathrm{\Lambda }0`$. In terms of the observable quantity $`H(z)\dot{a}(t)/a(t)`$ describing the evolution of the expanding Universe at recent epochs, the following inequality should be satisfied $$\frac{dH^2(z)}{dz}3\mathrm{\Omega }_{m,0}H_0^2(1+z)^2.$$ (1) Here, $`H_0=H(z=0)`$ is the Hubble constant, $`\mathrm{\Omega }_{m,0}`$ is the present energy density of the dustlike (CDM+baryons) matter component in terms of the critical density $`\epsilon _{\mathrm{crit}}=3H_0^2/8\pi G`$ ($`c=\mathrm{}=1`$, and an index $`0`$ stands for the present value of the corresponding quantity). Note that the inequality (1) saturates when the $`\mathrm{\Lambda }`$-term is exactly constant. It is not clear from the existing data whether (1) is satisfied at all. Actually the opposite holds: An attempt to reconstruct $`U(\mathrm{\Phi })`$ from the supernovae data and fitting of existing data to a model with a linear equation of state for the $`\mathrm{\Lambda }`$-term $`p_\mathrm{\Lambda }=w\epsilon _\mathrm{\Lambda }`$, with $`w<1`$ , shows that the possibility of violation of inequality (1), though strongly restricted, is not completely excluded. Hence it is natural and important to consider a variable $`\mathrm{\Lambda }`$-term in a more general class of scalar-tensor theories of gravity where the requirement (1) does not arise. Moreover, this generalization of general relativity (GR) is inspired by present more fundamental quantum theories, like $`M`$-theory. In these theories, the scalar field $`\mathrm{\Phi }`$ is just the dilaton field, hence we shall call it so below. Thus, we are interested in a universe where gravity is described by a scalar-tensor theory, and we consider the Lagrangian density in the Jordan frame $$L=\frac{1}{2}\left(F(\mathrm{\Phi })Rg^{\mu \nu }_\mu \mathrm{\Phi }_\nu \mathrm{\Phi }\right)U(\mathrm{\Phi })+L_m(g_{\mu \nu }),$$ (2) where $`L_m`$ describes dustlike matter and $`F(\mathrm{\Phi })>0`$. This corresponds to the Brans-Dicke parameter $`\omega =F/(dF/d\mathrm{\Phi })^2>0`$. One may also introduce a function $`Z(\mathrm{\Phi })`$ in front of the kinetic term $`(_\mu \mathrm{\Phi })^2`$, but it can be set either to $`1`$, or to $`1`$ by a redefinition of the scalar field. Under the assumption of absence of ghosts in the theory, the second possibility requires the Brans-Dicke parameter to lie in the range $`3/2<\omega <0`$ (see for more details). Since this clearly contradicts solar system tests of GR either in the absence of $`U(\mathrm{\Phi })`$, or for $`U(\mathrm{\Phi })`$ satisfying the condition (9) below for scales of galaxies and clusters of galaxies, we will not discuss this possibility further. We do not introduce any direct coupling between $`\mathrm{\Phi }`$ and $`L_m`$ (though this possibility could be envisaged, too). This guarantees that the weak equivalence principle is exactly satisfied (universality of free-fall of laboratory-size objects), and also that fundamental constants, like e.g. the fine-structure constant, do not change with time in this theory. This is in very good agreement with laboratory, geophysical and cosmological data . Such a scalar-tensor theory was recently considered as a model for a variable $`\mathrm{\Lambda }`$-term for some special choices of $`F(\mathrm{\Phi })`$ and $`U(\mathrm{\Phi })`$ (see ). Our approach is just the opposite: We want to derive these functions from observational data. Since we have to determine two functions $`F(\mathrm{\Phi })`$ and $`U(\mathrm{\Phi })`$, we will need both observational functions $`D_L(z)`$ and $`\delta _m(z)`$, in contrast to GR. Then the reconstruction problem can be uniquely solved as will be shown below. Note that the angular diameter as a function of $`z`$ provides the same information as $`D_L(z)`$ (see and the second reference in ). It is most appropriate for us to work in the Jordan frame (JF), in which the various physical quantities are those that are being measured in experiments, even though the Einstein frame (EF) often provides a better mathematical insight. In addition, the dilaton appears to be directly coupled to dustlike matter in the EF frame, in contrast to the JF. For a flat FRW universe with $`ds^2=dt^2+a^2d𝐱^2`$, the background equations in the JF are then $`3FH^2`$ $`=`$ $`\rho _m+{\displaystyle \frac{\dot{\mathrm{\Phi }}^2}{2}}+U3H\dot{F},`$ (3) $`2F\dot{H}`$ $`=`$ $`\rho _m+\dot{\mathrm{\Phi }}^2+\ddot{F}H\dot{F}.`$ (4) Their consequence is the equation for the dilaton itself: $$\ddot{\mathrm{\Phi }}+3H\dot{\mathrm{\Phi }}+\frac{dU}{d\mathrm{\Phi }}3(\dot{H}+2H^2)\frac{dF}{d\mathrm{\Phi }}=0.$$ (5) Combining Eqs. (3)–(4) and changing the argument from time $`t`$ to redshift $`z`$, we obtain the following basic equation for $`F(z)`$: $`F^{\prime \prime }`$ $`+`$ $`\left[(\mathrm{ln}H)^{}{\displaystyle \frac{4}{1+z}}\right]F^{}+\left[{\displaystyle \frac{6}{(1+z)^2}}{\displaystyle \frac{2(\mathrm{ln}H)^{}}{1+z}}\right]F`$ (6) $`=`$ $`{\displaystyle \frac{2U}{(1+z)^2H^2}}+3(1+z)\left({\displaystyle \frac{H_0}{H}}\right)^2F_0\mathrm{\Omega }_{m,0},`$ (7) where the prime denotes the derivative with respect to $`z`$. The effective value of Newton’s gravitational constant $`G_N`$ in Eqs. (34) is given by the formula $`G_N=1/8\pi F`$. We shall use its present value $`G_{N,0}`$ in the definition of the critical density $`\epsilon _{\mathrm{crit}}`$. On the other hand, $`G_{N,0}`$ is not the quantity measured in laboratory Cavendish-type and solar-system experiments. For a massless dilaton, the effective gravitational constant between two test masses is given by $$G_{\mathrm{eff}}=\frac{1}{8\pi F}\left(\frac{2F+4(dF/d\mathrm{\Phi })^2}{2F+3(dF/d\mathrm{\Phi })^2}\right).$$ (8) In our case, the dilaton is massive, so the expression (8) will be valid for physical scales $`R`$ such that $$R^2\mathrm{max}(\left|\frac{d^2U}{d\mathrm{\Phi }^2}\right|,H^2,H^2\left|\frac{d^2F}{d\mathrm{\Phi }^2}\right|).$$ (9) Previously, the expression $`G_{\mathrm{eff}}`$ was known from the post-Newtonian expansion; below we rederive it using the cosmological perturbation theory. Let us now list the restrictions of the theory (2) which follow from solar-system and cosmological tests. The post-Newtonian parameters $`\beta `$ and $`\gamma `$ for this theory are: $`\gamma `$ $`=`$ $`1{\displaystyle \frac{(dF/d\mathrm{\Phi })^2}{F+2(dF/d\mathrm{\Phi })^2}},`$ (10) $`\beta `$ $`=`$ $`1+{\displaystyle \frac{1}{4}}{\displaystyle \frac{F(dF/d\mathrm{\Phi })}{2F+3(dF/d\mathrm{\Phi })^2}}{\displaystyle \frac{d\gamma }{d\mathrm{\Phi }}}.`$ (11) Using the upper bounds on $`(\gamma 1)`$ from solar system measurements , we get $$\omega _0^1=F_0^1(dF/d\mathrm{\Phi })_0^2<4\times 10^4.$$ (12) So, $`G_{N,0}`$ and $`G_{\mathrm{eff},0}`$ coincide with better than $`2\times 10^4`$ accuracy. On the other hand, the difference between $`G_N`$ and $`G_{\mathrm{eff}}`$ may be larger at redshifts $`z1`$ since neither the upper limit on $`\beta `$, nor the present experimental bound $`|\dot{G}_{\mathrm{eff}}/G_{\mathrm{eff}}|<6\times 10^{12}\mathrm{yr}^1`$ significantly restrict $`(d^2F/d\mathrm{\Phi }^2)_0`$. Note that we cannot use the nucleosynthesis bound on the change of $`G_{\mathrm{eff}}`$ since that time as the behavior of $`G_{\mathrm{eff}}`$ during the intermediate period is unknown, unless we make additional assumptions (see below). The theory (2) describes a variable $`\mathrm{\Lambda }`$-term with desired properties if the following three conditions are satisfied: 1) The $`\mathrm{\Lambda }`$-term is dynamically important at present, namely, $`\mathrm{\Omega }_{\mathrm{\Lambda },0}0.72\mathrm{\Omega }_{m,0}`$, or $$\left(\frac{\dot{\mathrm{\Phi }}^2}{2}+U3H\dot{F}\right)_00.7\epsilon _{\mathrm{crit}}2\rho _{m,0}.$$ (13) 2) The $`\mathrm{\Lambda }`$-term has a sufficiently large negative pressure to provide acceleration of the present Universe. The condition $`\ddot{a}_0>0`$ reads: $$2U_0>(\rho _m+2\dot{\mathrm{\Phi }}^2+3\ddot{F}+3H\dot{F})_0.$$ (14) 3) The dark matter described by the $`\mathrm{\Lambda }`$-term remains unclustered at scales up to $`R10h^1(1+z)^1`$ Mpc and probably even more (here $`h=H_0/100`$ km s<sup>-1</sup> Mpc<sup>-1</sup>). To achieve this, it is sufficient to assume that the inequality (9) is satisfied for all scales in question. The first step of our program is purely kinematical: we determine $`H(z)`$ from $`D_L(z)`$ like in GR, $$\frac{1}{H(z)}=\left(\frac{D_L(z)}{1+z}\right)^{}.$$ (15) The functional dependence of $`D_L(z)`$ on the cosmological parameters, like $`\mathrm{\Omega }_{m,0}`$, is of course model dependent. If $`\mathrm{\Omega }_{m,0}`$ is already known from other tests, we can find already at that stage of the reconstruction a quantity such as the present effective equation of state of the dilaton from the formula (cf. ): $$w_0\frac{p_{\mathrm{\Lambda },0}}{\epsilon _{\mathrm{\Lambda },0}}=\frac{(2/3)(d\mathrm{ln}H/dz)_01}{1\mathrm{\Omega }_{m,0}}.$$ (16) $`\epsilon _{\mathrm{\Lambda },0}`$ contains the term $`3H_0\dot{F}_0`$, so that $`\mathrm{\Omega }_{m,0}+\mathrm{\Omega }_{\mathrm{\Lambda },0}=1`$. The dilaton equation of state can be determined for $`z>0`$, too; one has only to define what should be called the pressure and the energy density of the dilaton in general. Actually, we will show below that $`\mathrm{\Omega }_{m,0}`$ is itself self-consistently determined from our approach, so no additional information is required to find $`w(z)`$. In contrast to GR, Eq. (7) is no longer sufficient to determine $`U(z)`$; one should know $`F(z)`$, too. For this purpose we will use $`\delta _m(z)`$. We consider perturbations in the longitudinal gauge $`ds^2=(1+2\varphi )dt^2+a^2(12\psi )d𝐱^2`$. Working in Fourier space (a spatial dependence $`\mathrm{exp}(i𝐤.𝐱)`$ with $`k|𝐤|`$ is assumed), the following equations are obtained: $`\varphi `$ $`=`$ $`\dot{v}=\psi \delta F/F,`$ (17) $`\dot{\delta }_m`$ $`=`$ $`{\displaystyle \frac{k^2}{a^2}}v+3{\displaystyle \frac{d(\psi +Hv)}{dt}},`$ (18) where the gauge invariant quantity $`\delta _m(\delta \rho _m)/\rho _m+3Hv`$, and $`v`$ is the peculiar velocity potential of dustlike matter. We also get $`3\dot{F}\dot{\varphi }`$ $``$ $`\left(2{\displaystyle \frac{k^2}{a^2}}F\dot{\mathrm{\Phi }}^2+3H\dot{F}\right)\varphi =`$ (19) $`=`$ $`\rho _m\delta _m+3{\displaystyle \frac{\dot{F}}{F}}\dot{\delta F}+\left({\displaystyle \frac{k^2}{a^2}}6H^23{\displaystyle \frac{\dot{F}^2}{F^2}}\right)\delta F`$ (20) $`+`$ $`\dot{\mathrm{\Phi }}\dot{\delta \mathrm{\Phi }}+3H\dot{\mathrm{\Phi }}\delta \mathrm{\Phi }+\delta U,`$ (21) and the equation for the dilaton fluctuations $`\delta \mathrm{\Phi }`$: $`\ddot{\delta \mathrm{\Phi }}+3H\dot{\delta \mathrm{\Phi }}`$ $`+`$ $`\left[{\displaystyle \frac{k^2}{a^2}}3(\dot{H}+2H^2){\displaystyle \frac{d^2F}{d\mathrm{\Phi }^2}}+{\displaystyle \frac{d^2U}{d\mathrm{\Phi }^2}}\right]\delta \mathrm{\Phi }=`$ (22) $`=`$ $`\left[{\displaystyle \frac{k^2}{a^2}}(\varphi 2\psi )3(\ddot{\psi }+4H\dot{\psi }+H\dot{\varphi })\right]{\displaystyle \frac{dF}{d\mathrm{\Phi }}}`$ (24) $`+(3\dot{\psi }+\dot{\varphi })\dot{\mathrm{\Phi }}2\varphi {\displaystyle \frac{dU}{d\mathrm{\Phi }}}.`$ Let us now consider sufficiently small scales $`R=2\pi a/k`$ for which the inequality (9) is well satisfied. For example, if $`\delta _m(z)`$ is determined from the abundance of rich clusters of galaxies, then the relevant comoving scale is $`R8h^1/(1+z)`$ Mpc. If the r.h.s. of Eq. (9) is $`H^2`$, then the corresponding small parameter is $`R^2H_0^210^5`$. Note that we have another parameter, $`\omega ^1`$, which is small at the present time, Eq. (12), but it need not be so small in the past. Also, this parameter may be larger than $`a^2H^2/k^2`$. For this reason, we will first keep it. The solution of Eqs. (1724) in the formal short-wavelength limit $`k\mathrm{}`$ can be found following the analytical method used in in the GR case, confirmed numerically in . The idea is that the leading terms in Eqs. (1724) are either those containing $`k^2`$, or those with $`\delta _m`$. Then, using (18) and the l.h.s. of Eq. (17), the standard form of the equation for dustlike matter density perturbation follows: $$\ddot{\delta }_m+2H\dot{\delta }_m+k^2a^2\varphi 0.$$ (25) Now we consider the solution of Eq. (24) of interest to us, for which $`|\ddot{\delta \mathrm{\Phi }}|k^2a^2|\delta \mathrm{\Phi }|`$. It corresponds to the growing adiabatic mode. So, keeping terms with $`k^2`$ in Eq. (24) and then using the r.h.s. of Eq. (17), we obtain: $$\delta \mathrm{\Phi }(\varphi 2\psi )\frac{dF}{d\mathrm{\Phi }}\varphi \frac{FdF/d\mathrm{\Phi }}{F+2(dF/d\mathrm{\Phi })^2}.$$ (26) In the GR case, $`\delta \mathrm{\Phi }k^2\varphi `$ in the limit $`k\mathrm{}`$, so matter producing the $`\mathrm{\Lambda }`$-term is not gravitationally clustered at small scales (physically, due to free streaming). This is not so in scalar-tensor gravity: The dilaton remains partly clustered for arbitrarily small scales, this clustering being small only because $`\omega `$ is large. Keeping only terms with $`k^2`$ or $`\delta _m`$ in Eq. (21), we get the expression of $`\varphi `$ through $`\delta _m`$ and $`\delta F`$. Finally, inserting it into Eq. (25) and using Eq. (26), we arrive to the closed form of the equation for $`\delta _m`$: $$\ddot{\delta }_m+2H\dot{\delta }_m4\pi G_{\mathrm{eff}}\rho _m\delta _m0,$$ (27) with $`G_{\mathrm{eff}}`$ defined in (8) above. In terms of $`z`$, (27) reads: $`H^2\delta _m^{\prime \prime }+\left({\displaystyle \frac{(H^2)^{}}{2}}{\displaystyle \frac{H^2}{1+z}}\right)\delta _m^{}`$ (28) $`{\displaystyle \frac{3}{2}}(1+z)H_0^2{\displaystyle \frac{G_{\mathrm{eff}}(z)}{G_{N,0}}}\mathrm{\Omega }_{m,0}\delta _m.`$ (29) Eq. (27) does not contain $`k^2`$ at all. Thus, its solutions, as well as the corresponding expressions for $`\delta \mathrm{\Phi }`$, do not oscillate with the frequency $`k/a`$ for $`k\mathrm{}`$. This justifies the assumption about $`\ddot{\delta \mathrm{\Phi }}`$ made above. Extracting $`H(z)`$ (from $`D_L(z)`$) and $`\delta _m(z)`$ from observations with sufficient accuracy, we can reconstruct $`G_{\mathrm{eff}}(z)/G_{N,0}`$ analytically. Since, as follows from Eq. (12), the quantities $`G_{\mathrm{eff},0}`$ and $`G_{N,0}`$ coincide with better than $`0.02\%`$ accuracy, Eq. (29) taken at $`z=0`$ gives also the value of $`\mathrm{\Omega }_{m,0}`$ with the same accuracy. Thus, in principle, no independent measurement of $`\mathrm{\Omega }_{m,0}`$ is required. The resulting equation $`G_{\mathrm{eff}}(z)=p(z)`$, where $`p(z)`$ is a given function following from observational data, can be transformed into a nonlinear second order differential equation for $`F(z)`$ if we exclude $`d\mathrm{\Phi }`$ (which appears in $`dF/d\mathrm{\Phi }`$) using the background equation (4), which reads $`\mathrm{\Phi }^2`$ $`=`$ $`F^{\prime \prime }\left[(\mathrm{ln}H)^{}+{\displaystyle \frac{2}{1+z}}\right]F^{}`$ (31) $`+{\displaystyle \frac{2(\mathrm{ln}H)^{}}{1+z}}F3(1+z){\displaystyle \frac{H_0^2}{H^2}}F_0\mathrm{\Omega }_{m,0}.`$ Therefore, $`F(z)`$ can be determined by solving that equation provided $`F_0`$ ($`=1/8\pi G_{N,0})`$ and $`F_0^{}`$ are known. However, this procedure can be simplified a lot under reasonable assumptions, and taking into account the small present values of $`\omega ^1=F^1(dF/d\mathrm{\Phi })^2`$ and $`\dot{G}_{\mathrm{eff}}/G_{\mathrm{eff}}`$. Indeed, the value of $`\omega ^1`$ for $`0z1`$ can be estimated from the first terms of its Taylor expansion $`\omega _0^1+z(d\omega ^1/dz)_0`$. Neglecting contributions proportional to $`\omega _0^1`$, we then get $`\omega ^12z\lambda (d^2F/d\mathrm{\Phi }^2)_0`$, with $`\lambda (d\mathrm{ln}F/d\mathrm{\Phi })_0\dot{\mathrm{\Phi }}_0/H_0`$, whereas $`\dot{G}_{\mathrm{eff}}/G_{\mathrm{eff}}\lambda H_0[1(d^2F/d\mathrm{\Phi }^2)_0]`$. If $`(d^2F/d\mathrm{\Phi }^2)_0`$ differs significantly from 1, we can thus conclude that $`\omega ^1|2\dot{G}_{\mathrm{eff}}/H_0G_{\mathrm{eff}}|0.25`$. On the other hand, if $`(d^2F/d\mathrm{\Phi }^2)_0`$ happens to be close to 1, one can still assume that there is no special cancellation of large terms in the r.h.s. of Eq. (3), and therefore that $`\dot{\mathrm{\Phi }}_0^26F_0H_0^2`$. The above estimate for $`\omega ^1`$ then gives $`\omega ^12\sqrt{6/\omega _0}0.1`$. In both cases, we thus find that $`G_{\mathrm{eff}}G_N`$ in the range of $`z`$ involved with better than $`10\%`$ accuracy. Note that the same estimate may be obtained by assuming that $`\omega ^1`$ changed monotonically with $`z`$ and using the nucleosynthesis bound (cf. ). Therefore, in first approximation in $`\omega ^1`$, $`G_{\mathrm{eff}}(z)1/8\pi F(z)`$ and Eq. (29) can be used to determine $`F(z)`$ unambiguously. Small corrections to this result can be taken into account using perturbation theory with respect to the small parameter $`\omega ^1`$. After $`F(z)`$ is found, the potential $`U(z)`$ is determined from Eq. (7). Finally, using Eq. (31) we find $`\mathrm{\Phi }(z)`$ by simple integration. After that, both unknown functions $`F(\mathrm{\Phi })`$ and $`U(\mathrm{\Phi })`$ are completely fixed as functions of $`\mathrm{\Phi }\mathrm{\Phi }_0`$ in that range probed by the data. Equations (29), (7) and (31), giving the subsequent steps of the reconstruction, constitute the fundamental result of our letter. Our results generalize those obtained in GR and constrain any attempt to explain a varying $`\mathrm{\Lambda }`$-term using scalar-tensor theories of gravity. Good data on $`\delta _m(z)`$ expected to appear soon from observations of clustering and abundance of different objects at redshifts $`1`$ and more, as well as from weak gravitational lensing, together with better data on $`D_L(z)`$ from more supernova events, will allow implementation of the reconstruction program and determination of the microscopic Lagrangian. ###### Acknowledgements. A.S. was partially supported by the Russian Foundation for Basic Research, grant 99-02-16224, and by the Russian Research Project “Cosmomicrophysics”. Centre de Physique Théorique is Unité Propre de Recherche 7061.
warning/0001/math0001130.html
ar5iv
text
# Veronese webs for bihamiltonian structures of higher corank ## 0 Introduction. A $`C^{\mathrm{}}`$\- manifold $`M`$ is endowed by a Poisson pair if two linearly independent smooth bivectors $`c_1,c_2`$ are defined on $`M`$ and $`c_\lambda =\lambda _1c_1+\lambda _2c_2`$ is a Poisson bivector for any $`\lambda =(\lambda _1,\lambda _2)R^2`$. A bihamiltonian structure $`J=\{c_\lambda \}`$ is the whole 2-dimensional family of bivectors. The structure $`J`$ is degenerate if $`rankc_\lambda <dimM,\lambda R^2`$. An intensive study of such objects was done by I.M.Gelfand and I.S.Zakharevich (, , ) in a particular case of bihamiltonian structures in general position on an odd-dimensional $`M`$ (the corresponding Poisson pairs are necessarily degenerate: $`rankc_\lambda =2n,\lambda R^2\{0\}`$, if $`dimM=2n+1`$). In there was introduced a notion of a Veronese web, i.e. a 1-parameter family of 1-codimensional foliations such that the corresponding family of annihilators is represented by the Veronese curve in the cotangent space at each point. It turns out that Veronese webs form a complete system of local invariants for bihamiltonian structures of general position. More precisely, it was shown in that any such structure $`J=\{c_\lambda \}`$ in $`R^{2n+1}`$ admits a local reduction to a Veronese web $`𝒲_J`$ on a $`(n+1)`$-dimensional manifold and that for any Veronese web $`𝒲`$ one can locally construct a bihamiltonian structure $`J(𝒲)`$ of general position in $`R^{2n+1}`$ with the reduction equal to $`𝒲`$. In the real analytic case $`J`$ and $`J(𝒲_J)`$ are isomorphic. The aim of this paper is to introduce a wider class of degenerate bihamiltonian structures that possess many features of the general position case and to generalize the notion of a Veronese web for this class. We call the bihamiltonian structures from this class complete since they are intemately connected with the completely integrable systems () on $`M`$. In particular, the Poisson pairs appearing in the well known method of argument translation (see , , and Example 1.11. Definition, below) generate complete bihamiltonian structures of higher ($`>1`$) corank. The paper is organized as follows. In Section 1 we recall some definitions and facts about bihamiltonian structures and introduce the main definition of completeness. The last is based on one result of A.Brailov (Theorem 1). We show that complete bihamiltonian structures generalize the case of general position. Analyzing the corresponding Poisson pair $`(c_1(x),c_2(x))`$ at a point $`xM`$ we deduce that it consists of finite number of the so called Kronecker blocks (Corollary 1.14. Theorem); the general position is characterized by the case of the sole block. Section 2 is devoted to distinguishing the invariants for the sum of $`k`$ Kronecker blocks. In the next section we define local Veronese webs for complete bihamiltonian structures under some additional assumption of simplicity. This last means that: 1)the number of Kronecker blocks does not change from point to point and the corresponding subspaces vary smoothly ”sweeping” a flag of $`k`$ subbundles in the tangent bundle; 2)there are no blocks of equal dimension. The second condition allows to avoid some technical complications but in principle may be skiped (see Remark 2). In general, the mentioned distributions are nonintegrable (Examples 3, 3.4. Example); consequently, the bihamiltonian structure does not split to direct product of the bihamiltonian structures of corank $`1`$, i.e. of general position. We conclude the paper calculating the Veronese web for the method of argument translation (Section 5). In the case of normal noncompact real form of complex simple Lie algebra this web is generically a product of flat Veronese webs of codimension $`1`$. Recent papers , are closely related to the subject, in particular to generalized Veronese webs. In the author introduces a more general notion of a Kronecker web, which is essentially equivalent to the notion of a Veronese web (see Definition 3.1. Definition) in case of simple bihamiltonian structures. Our approach emphasizes a bit more the role of Veronese curves in the theory. The following two questions arise from the context of this paper. 1. Does the Veronese web of a complete bihamiltonian structure determine it up to an isomorphism? 2. What is a relation between the Veronese webs introduced here and $`d`$-webs of maximal rank and codimension $`2`$ studied in paper of S.S.Chern and P.A.Griffiths? (The notion of the rank of a $`d`$-web should not be confused with that of a bihamiltonian structure; corank of bihamiltonian structure is equal to codimension of the web.) Note that the $`d`$-webs of maximal rank and codimension $`1`$ considered in paper of the same authors are intemately connected with the Veronese webs of codimension $`1`$. The author would like to thank Prof. Ilya Zakharevich for useful remarks on this paper and for indicating references , which had an essential inluence on its final version. ## 1 Bihamiltonian structures and completeness. Let $`M`$ be a $`C^{\mathrm{}}`$\- manifold. In the sequel, all considered Poisson bivectors will have maximal rank on an open dense subset in $`M`$. Given a Poisson bivector $`c`$, define $`rankc`$ as $`\mathrm{max}_{xM}rankc(x)`$. ###### 1.1. Definition Two linearly independent Poisson bivectors $`c_1,c_2`$ on $`M`$ form a Poisson pair if $`c_\lambda =\lambda _1c_1+\lambda _2c_2`$ is a Poisson bivector for any $`\lambda =(\lambda _1,\lambda _2)R^2`$. ###### 1.2. Proposition A pair of linearly independent Poisson bivectors $`(c_1,c_2)`$ is Poisson if and only if $`[c_1,c_2]=0`$, where $`[,]`$ is the Schouten bracket. ###### 1.3. Definition A bihamiltonian structure on $`M`$ is defined as a two-dimensional linear subspace $`J=\{c_\lambda \}_{\lambda 𝒮}`$ of Poisson bivectors on $`M`$ parametrized by a two-dimensional vector space $`𝒮`$ over $`R`$ . We say that $`J`$ is degenerate if $`rankc_\lambda <dimM`$ for any $`c_\lambda J`$. It is clear that every Poisson pair generates a bihamiltonian structure and the transition from the latter one to a Poisson pair corresponds to a choice of basis in $`𝒮`$. We shall write $`(J,c_1,c_2)`$ for a bihamiltonian structure $`J`$ with a chosen Poisson pair $`(c_1,c_2)`$ generating $`J`$. ###### 1.4. Definition Let $`J`$ be a bihamiltonian structure. Introduce a subfamily $`J_0J`$ of Poisson bivectors of maximal rank $`R_0`$ (the set $`JJ_0`$ is at most a finite sum of 1-dimensional subspaces), and a set of functions $`_0=Span_R(_{cJ_0}Z_c(M))`$, where $`Z_c(M)`$ stands for the space of the Casimir functions of $`c`$ on $`M`$. We take $`Span`$ in order to obtain a vector space: a sum of two Casimir functions for different $`c_1,c_2J_0`$ need not be a Casimir function. The following proposition shows how the degenerate bihamiltonian structures can be applied for constructing the completely integrable systems. ###### 1.5. Proposition Let $`J`$ be a degenerate bihamiltonian structure on $`M`$. A family $`_0`$ is involutive with respect to any $`c_\lambda J`$. Proof. Let $`c_1,c_2J_0`$ be linearly independent, $`f_iZ_{c_i},i=1,2`$. Then $$\{f_1,f_2\}_{c_\lambda }=(\lambda _1c_1(f_1)+\lambda _2c_2(f_1))f_2=\lambda _2c_2(f_2)f_1=0.$$ (1.5.1) Now it remains to prove that for any $`cJ_0,f_iZ_c,i=1,2`$, one has $`\{f_1,f_2\}_{c_\lambda }=0`$. For that purpose we first rewrite (1.5. Proposition) as $$c_\lambda (x)(\varphi _1,\varphi _2)=0,$$ (1.5.2) where $`\varphi _ikerc_i(x),i=1,2,xM`$, and the lefthandside denotes the contraction of the bivector with two covectors. Second, we fix $`x`$ such that $`rankc(x)=R_0`$ and approximate $`df_2|_x`$ by a sequence of elements $`\{\varphi ^i\}_{i=1}^{\mathrm{}},\varphi ^ikerc^i(x)`$, where $`c^iJ_0,i=1,2,\mathrm{},`$ is linearly independent with $`c`$. Finally, by (1.5.1) we get $`c_\lambda (x)(df_1|_x,\varphi ^i)=0`$ and by the continuity $`\{f_1,f_2\}_{c_\lambda }(x)=0`$. Since the set of such points $`x`$ is dense in $`M`$, the proof is finished. In fact this proposition is true for the local Casimir functions (for the germs of Casimir functions). The corresponding family of functions (germs) $`Span_R(_{cJ_0}Z_c(U))`$ ($`Span_R(_{cJ_0}Z_{c,x}`$) is denoted by $`_0(U)`$ ($`_{0,x}`$). In order to obtain a completely integrable system from Casimir functions one should require additional assumptions on the bihamiltonian structure $`J`$. Off course, the condition of completeness given below concerns the local Casimir functions (in fact their germs) and may be insufficient for obtaining the completely integrable system. However, it is of use if the local Casimir functions are restrictions of the global ones (see Example 1.11. Definition, below). Given a Poisson bivector $`c_\lambda J`$, let $`S_\lambda (x)`$ denote the symplectic leaf of $`c_\lambda `$ through a point $`xM`$. ###### 1.6. Definition () Let $`J`$ be a bihamiltonian structure; fix some $`c_\lambda J`$. $`J`$ is called complete at a point $`xM`$ with respect to $`c_\lambda `$ if the linear subspace of $`T_x^{}M`$ generated by the differentials of the germs $`f_{0,x}`$ restricted to $`S_\lambda (x)`$ has dimension $`\frac{1}{2}dimS_\lambda (x)`$. ###### 1.7. Proposition A bihamiltonian structure $`J`$ is complete with respect to $`c_\lambda J_0`$ at a point $`xM`$ such that $`S_\lambda (x)`$ is of maximal dimension if and only if $`dim(_{c_\lambda J_0}T_xS_\lambda (x))=\frac{1}{2}dimS_\lambda (x)`$. The following theorem is due to A.Brailov (see , Theorem 1.1 and Remark after it). ###### 1.8. Theorem A bihamiltonian structure $`(J,c_1,c_2)`$ is complete with respect to $`c_\lambda J_0`$ at a point $`xM`$ such that $`S_\lambda (x)`$ is of maximal dimension if and only if the following condition holds $`rank(\lambda _1c_1+\lambda _2c_2)(x)=R_0`$ for any $`\lambda =(\lambda _1,\lambda _2)C^2\{0\}`$. Here the bivector $`c_\lambda =(\lambda _1c_1+\lambda _2c_2)(x)`$ is regarded as an element of $`^2T_x^CM`$, where $`T^CM`$ is the complexified tangent bundle, and its rank is defined as that of the associated sharp map $`c_\lambda ^{\mathrm{}}(x):(T_x^CM)^{}T_x^CM`$. The theorem shows that $`J`$ is complete with respect to a fixed $`c_\lambda J_0`$ at a point $`x`$ such that the dimension $`S_\lambda (x)`$ is maximal if and only if $`J=J_0\{0\}`$ and $`J`$ is complete at $`x`$ with respect to any nontrivial $`c_\lambda J`$. This motivates the next definition. ###### 1.9. Definition Let $`(J,c_1,c_2)`$ be a bihamiltonian structure. The structure $`J`$ (the pair $`(c_1,c_2)`$) is complete at a point $`xM`$ if condition $`()`$ of Theorem 1 holds at $`x`$. $`J`$ ($`(c_1,c_2)`$) is called complete if it is so at any point from some open and dense subset in $`M`$. ###### 1.10. Proposition Let $`J`$ be complete on $`M`$ and let $`xM`$ be a point of completeness. Then there exists a neighbourhood $`Ux`$ such that the foliation $``$ defined on $`U`$ by $`_0(U)`$ is lagrangian in any $`S_\lambda (y),\lambda 0,yU`$ (by Proposition 1.6. Definition this foliation can be defined as the intersection of the foliations of symplectic leaves for $`c_\lambda J_0`$). ###### 1.11. Definition Call $``$ a bilagrangian foliation of $`J`$. ###### 1.12. Example (Method of argument translation, see , .) Let $`g`$ be a Lie algebra, $`g^{}`$ its dual space. Fix a basis $`\{e_1,\mathrm{},e_n\}`$ in $`g`$ with the structure constants $`\{c_{ij}^k\}`$; write $`\{e^1,\mathrm{},e^n\}`$ for the dual basis in $`g^{}`$. The standard linear Poisson bivector on $`g^{}`$ is defined as $$c_1(x)=c_{ij}^kx_k\frac{}{x_i}\frac{}{x_j},$$ where $`\{x_k\}`$ are linear coordinates in $`g^{}`$ corresponding to $`\{e^1,\mathrm{},e^n\}`$. In more invariant terms $`c_1`$ is described as an operator dual to the Lie-multiplication map $`[,]:ggg`$. It is well-known that the symplectic leaves of $`c_1`$ are the coadjoint orbits in $`g^{}`$. Now define $`c_2`$ as a bivector with constant coefficients $`c_2=c(a)`$, where $`a`$ is a fixed point on any leaf of maximal dimension. It turns out that $`c_1,c_2`$ form a Poisson pair and it is easy to describe the set $`I`$ of points $`x`$ for which condition $`()`$ fails. Consider the complexification $`(g^{})^C(g^𝐂)^{}`$ and the sum $`Sing(g^C)^{}`$ of symplectic leaves of nonmaximal dimension for the complex linear bivector $`c_{ij}^kz_k\frac{}{z_i}\frac{}{z_j},`$ where $`z_j=x_j+\mathrm{i}y_j,j=1,\mathrm{},n,`$ are the corresponding complex coordinates in $`(g^{})^C`$. Then $`I`$ is equal to the intersection of the sets $`g^{}(g^{})^C`$ and $`\overline{a,Sing(g^C)^{}}`$, where $`\overline{a,Sing(g^C)^{}}`$ denotes a cone of complex $`2`$-dimensional subspaces passing through $`a`$ and $`Sing(g^C)^{}`$. In particular, $`(c_1,c_2)`$ is complete for a semisimple $`g`$ ($`codimSing(g^C)^{}3`$, see , Corollary 4.42, and codimension of $`I`$ in $`g^{}`$ is not less than $`2`$). Note, that this gives rise to completely integrable systems since the local Casimir functions on $`g^{}`$ are restrictions of the global ones, i.e. the invariants of the coadjoint action. ###### 1.13. Example (Bihamiltonian structure of general position on an odd-dimensional manifold, see .) Consider a pair of bivectors $`(a_1,a_2)`$, $`a_i^2V,i=1,2`$, where $`V`$ is a $`(2m+1)`$-dimensional vector space; $`(a_1,a_2)`$ is in general position if and only if is represented by the Kronecker block of dimension $`2m+1`$, i.e. $$\begin{array}{c}a_1=p_1q_1+p_2q_2+\mathrm{}+p_mq_m\hfill \\ a_2=p_1q_2+p_2q_3+\mathrm{}+p_mq_{m+1}\hfill \end{array}$$ (1.13.0) in an appropriate basis $`p_1,\mathrm{}p_m,q_1,\mathrm{}q_{m+1}`$ of $`V`$. A bihamiltonian structure $`J`$ on a $`(2m+1)`$-dimensional $`M`$ is in general position if and only if the pair $`(c_1(x),c_2(x))`$ is so for any $`xM`$. Such $`J`$ is complete. In general, a complete Poisson pair at a point is a direct sum of the Kronecker blocks as the corollary of the next theorem shows. This theorem is a reformulation of the classification result for pairs of $`2`$-forms in a vector space (, ). ###### 1.14. Theorem Given a finite-dimensional vector space $`V`$ over $`C`$ and a pair of bivectros $`(c_1,c_2),c_i^2V,`$ there exists a direct decomposition $`V=V_j,c_i=c_i^{(j)},c_i^{(j)}^2V_j,i=1,2,`$ such that each triple $`(V_j,c_1^{(j)},c_2^{(j)})`$ is from the following list: the Jordan block: $`dimV_j=2n_j`$ and in an appropriate basis of $`V_j`$ the matrix of $`c_i^{(j)}`$ is equal to $$\left(\begin{array}{cc}0& A_i\\ A_i^T& 0\end{array}\right),i=1,2,$$ where $`A_1=I_{n_j}`$ (the unity $`n_j\times n_j`$-matrix) and $`A_2=J_{n_j}^\lambda `$ (the Jordan block with the eigenvalue $`\lambda `$); the Kronecker block: $`dimV_j=2n_j+1`$ and in an appropriate basis of $`V_j`$ the matrix of $`c_i^{(j)}`$ is equal to $$\left(\begin{array}{cc}0& B_i\\ B_i^T& 0\end{array}\right),i=1,2,$$ where $`B_1=\left(\begin{array}{cccccc}1& 0& 0& \mathrm{}& 0& 0\\ 0& 1& 0& \mathrm{}& 0& 0\\ & & & \mathrm{}& & \\ 0& 0& 0& \mathrm{}& 1& 0\end{array}\right),B_2=\left(\begin{array}{cccccc}0& 1& 0& \mathrm{}& 0& 0\\ 0& 0& 1& \mathrm{}& 0& 0\\ & & & \mathrm{}& & \\ 0& 0& 0& \mathrm{}& 0& 1\end{array}\right)`$ ($`(n_j+1)\times n_j`$-matrices; in case $`n_j=0`$ put $`c_1=c_2=0`$). ###### 1.15. Corollary Let $`(J,c_1,c_2)`$ be a bihamiltonian structure. It is complete at a point $`xM`$ if and only if the pair $`(c_1(x),c_2(x)),c_i(x)^2(T_x^CM),i=1,2,`$ does not contain the Jordan blocks in its decomposition. Proof. The statement follows from the definition of completeness. ###### 1.16. Remark In the absence of the Jordan blocks Theorem 1.13. Example is valid also over reals. ## 2 Complete bihamiltonian structure at a point. Now, we shall examine a linear bihamiltonian structure $`(J,c_1,c_2),c_i^2V`$, where $`V`$ is a vector space over $`R`$, such that the decomposition $`V=_{j=1}^kV_j,c_i=_{j=1}^kc_i^{(j)}`$ (see Theorem 1.13. Example and Remark 1) consists of $`k`$ Kronecker blocks $`V_1,\mathrm{},V_k,dimV_j=2n_j+1,n_1<\mathrm{}<n_k`$. The aim is to extract the invariants and to introduce the infinithesimal approximation to Veronese webs (these last will be defined in the next section). It turns out that the decomposition to Kronecker blocks is noninvariant. To illustrate this let us consider $`V=Span\{e,p,q_1,q_2\},c_1=pq_1,c_2=pq_2`$. Here $`V=V_1V_2`$, where $`V_1=Span\{e\},V_2=Span\{p,q_1,q_2\}`$, but instead $`V_1`$ one can choose any direct complement to $`V_2`$. However, there is a canonically defined filtration associated to $`J`$. Let $`P_{c_\lambda }V`$ ($`P_{c_\lambda ^{(j)}}V_j`$) be the characteristic subspace, i.e. the symplectic leaf through $`0`$, of $`c_\lambda =\lambda _1c_1+\lambda _2c_2`$ ($`c_\lambda ^{(j)}=\lambda _1c_1^{(j)}+\lambda _2c_2^{(j)}`$), $`(\lambda _1,\lambda _2)R^2`$; let $`L=_{\lambda 0}P_{c_\lambda }`$ ($`L_j=_{\lambda 0}P_{c_\lambda ^{(j)}}`$) be the bilagrangian subspase, i.e. the leaf through $`0`$ of the bilagrangian foliation, corresponding to the bihamiltonian structure $`J`$ ($`\{c_\lambda ^{(j)}\}_{\lambda R^2}`$), see Definition 1.10. Proposition. Put $$\mathrm{\Phi }_i=\underset{\text{distinct}\lambda _1,\mathrm{},\lambda _iP(R^2)}{}P_{c_{\lambda _1}}\mathrm{}P_{c_{\lambda _i}},i=1,2,\mathrm{},\mathrm{\Phi }_0=V.$$ ###### 2.1. Theorem The following relations hold $$\mathrm{\Phi }_0=\mathrm{\Phi }_1=\mathrm{}=\mathrm{\Phi }_{n_1}\mathrm{\Phi }_{n_1+1}=\mathrm{\Phi }_{n_1+2}=\mathrm{}=\mathrm{\Phi }_{n_2}\mathrm{}\mathrm{\Phi }_{n_{k1}+1}=\mathrm{}=\mathrm{\Phi }_{n_k}$$ $$\mathrm{\Phi }_{n_k+1}=\mathrm{\Phi }_{n_k+2}=\mathrm{}=:\mathrm{\Phi }_{n_{k+1}},$$ where $$\mathrm{\Phi }_{n_j}=\underset{l<j}{}L_l\underset{lj}{}V_l,j=1,\mathrm{},k+1$$ (we put $`V_l=0,l>k`$). In particular, the filtration is stabilized from $`i>n_k`$, $`\mathrm{\Phi }_{n_k+1}=L_j=L`$, and the numbers $`n_1,\mathrm{},n_k`$ are invariants of $`J`$. ###### 2.2. Remark The filtration $`F_0=\mathrm{\Phi }_{n_1}^{}\mathrm{}F_{k1}=\mathrm{\Phi }_{n_k}^{}F_k=L^{}=(V/L)^{}`$ ($``$ stands for the annihilator sign) appears in and is called there isotypic. We shall refer to this notion below. Before we begin to prove the theorem we recall the following definition. ###### 2.3. Definition () Let $`𝒮,V`$ be vector spaces of dimensions $`2`$ and $`n+1,n0`$, respectively. A Veronese inclusion of $`P(𝒮)`$ in $`P(V)`$ is a map $`i:P(𝒮)P(V)`$ such that there exists a linear isomorphism $`\varphi :P(V)P(S^n𝒮)`$ making the following diagram commutative: $$\begin{array}{ccc}P(𝒮)& \stackrel{i}{}& P(V)\\ & & \varphi \\ P(𝒮)& \stackrel{P(S^n())}{}& P(S^n𝒮).\end{array}$$ The image $`i(P(𝒮))`$ is called Veronese curve. Here $`S^n`$ denotes the $`n`$-th symmetric power; the standard model of the mapping $`S^n()`$ is described as follows. Let $`𝒮`$ be a space of linear functions $`f`$ in two variables $`t_1,t_2`$. Then $`S^n𝒮`$ is a space of homogeneous polynomials in $`t_1,t_2`$ and $`S^n(f)=f^n`$. For $`n=0`$ the map $`i`$ is not an inclusion, but we shall use the defined term in this situation as well. Proof. We now shall prove Theorem 2. It is sufficient to show the following equalities $$\mathrm{\Phi }_iV_j:=\underset{\text{distinct}\lambda _1,\mathrm{},\lambda _i}{}P_{c_{\lambda _1}^{(j)}}\mathrm{}P_{c_{\lambda _i}^{(j)}}=\{\begin{array}{cc}L_j\hfill & in_j+1\hfill \\ V_j\hfill & i<n_j+1.\hfill \end{array}$$ (2.3.1) One has $$(\mathrm{\Phi }_iV_j)^{}=\underset{\text{distinct}\lambda _1,\mathrm{},\lambda _i}{}(P_{c_{\lambda _1}^{(j)}}^{}+\mathrm{}+P_{c_{\lambda _i}^{(j)}}^{}).$$ The 1-dimentional annihilator $`P_{c_\lambda ^{(j)}}^{}P((V_j/L_j)^{})`$ sweeps an appropriate Veronese curve (see ). On the other hand, images of distinct points under a Veronese inclusion $`P(𝒮)P(V)`$ are linearly independent untill their number does not exceed $`dimV`$ (cf. , I.A). So now, the first of equalities 2.3. Definition follows from the fact that $`P_{c_{\lambda _1}^{(j)}}^{}+\mathrm{}+P_{c_{\lambda _i}^{(j)}}^{}=(V_j/L_j)^{}`$ for any set of distinct $`\lambda _1,\mathrm{},\lambda _i,in_j+1`$. For the second one, we notice that for any set of points $`\lambda _1,\mathrm{},\lambda _{n_j+1}`$ $$\underset{s=1}{\overset{n_j+1}{}}(P_1+\mathrm{}+\widehat{P_s}+\mathrm{}+P_{n_j+1})=\{0\},$$ where we put $`P_s=P_{c_{\lambda _s}}^{}`$ and $`\widehat{}`$ means omitting of the corresponding term. Thus we proved it for $`i=n_j`$; this implies 2.3. Definition also for $`i<n_j`$. ###### 2.4. Corollary Let $`0=F_1\mathrm{}F_{k1}F_k=(V/L)^{}`$ be the isotypic filtration (see 2.1. Theorem). Put $`F_jP_{c_\lambda }^{}=F_jP_{c_\lambda }^{},j=1,\mathrm{},k`$. Then: 1) $`A_j^\lambda :=F_jP_{c_\lambda }^{}/F_{j1}P_{c_\lambda }^{}`$ is a one-dimensional subspace in $`A_j:=F_j/F_{j1}`$; 2) the mapping $`P(R^2)\lambda A_j^\lambda P(A_j)`$ is a Veronese inclusion for any $`j=1,\mathrm{},k`$. Proof. We first notice that $`F_j=_{i=1}^j(V_i/L_i)^{},j=1,\mathrm{},k`$, as Theorem 2 implies. Under the identification $`(V/L)^{}=_{i=1}^k(V_i/L_i)^{}`$ $$F_jP_{c_\lambda }^{}=_{i=1}^jP_{c_\lambda ^{(i)}}^_i,$$ where $`_i`$ stands for the annihilator of a subspace in $`V_i/L_i`$. Thus there are linear isomorphisms $`A_j(V_j/L_j)^{}`$ and $`A_j^\lambda P_{c_\lambda ^{(j)}}^_j,j=1,\mathrm{},k`$, and 2) follows from the analogous fact for a sole Kronecker block (). ###### 2.5. Remark We note that analogues of Theorem 2 and Corollary 2 can be proved also without the restriction that $`n_1,\mathrm{},n_k`$ are distinct. In order to do that one should use a ”multiple” version of a Veronese inclusion, i.e. a map $`\varphi :P(𝒮)G(k,V^{(l+1)k})`$ ($`G(k,V)`$ denotes the Grassmannian of $`k`$-planes in a $`(l+1)k`$-dimensional vector space $`V`$) such that there exist a decomposition $`V=_{j=1}^kV_j,dimV_j=l+1`$, and Veronese inclusions $`i_j:P(𝒮)P(V_j),j=1,\mathrm{},k`$, with the property $`\varphi (v)=Span_R\{i_1(v),\mathrm{},i_k(v)\}`$, where $`i_j(v)`$ is considered as a 1-dimensional subspace in $`V`$. This definition can be also used to adapt the notion of a Veronese web (see Definition 3.1. Definition, below) to a more general situation. ###### 2.6. Definition An infinithesimal Veronese web of type $`(n_1,\mathrm{},n_k),n_1<\mathrm{}<n_k`$, on a vector space $`W,dimW=n_1+\mathrm{}+n_k+k`$, is a 1-parameter family $`\{𝒲_\lambda \}_{\lambda P(𝒮)}`$ of linear subspaces $`𝒲_\lambda W,codim𝒲_\lambda =k`$, satisfying the following conditions: there is a filtration $`0=F_0\mathrm{}F_{k1}F_k=W^{}`$ of the dual space with $`dimF_j/F_{j1}=n_j+1,j=1,\mathrm{},k`$; it induces the filtration $`0=F_0𝒲_\lambda ^{}\mathrm{}F_{k1}𝒲_\lambda ^{}F_k𝒲_\lambda ^{}=𝒲_\lambda ^{},F_j𝒲_\lambda ^{}=F_j𝒲_\lambda ^{},j=1,\mathrm{},k`$, of the annihilator $`𝒲_\lambda ^{}W^{}`$ so that $`dimF_j𝒲_\lambda ^{}=j`$; in particular $`A_j^\lambda :=F_j𝒲_\lambda ^{}/F_{j1}𝒲_\lambda ^{}`$ can be considered as 1-dimensional subspace in $`A_j:=F_j/F_{j1}`$; the map $`P(𝒮)\lambda A_j^\lambda P(A_j)`$ is a Veronese inclusion, $`j=1,\mathrm{},k`$. ###### 2.7. Proposition Let $`(J,c_1,c_2)`$ be as above. Then the vector space $`W=V/L`$ has a structure of an infinithesimal Veronese web of type $`(n_1,\mathrm{},n_k)`$. Proof. The proof follows from Corollary 2. ## 3 Simple bihamiltonian structures and their Veronese webs. In this section we shall define objects that generalize the Veronese webs introduced in for the bihamiltonian structures of general position. We shall show that any complete bihamiltonian structure from the class defined below admits the local reduction to such an object. ###### 3.1. Definition Let $`J`$ be a complete bihamiltonian structure on $`M`$. A type of $`J`$ at $`xM`$ is the vector $`(n_1,\mathrm{},n_k)(x)`$, where $`2n_1(x)+1,\mathrm{},2n_k(x)+1`$ are dimensions of the Kronecker blocks in the decomposition of $`(c_1(x),c_2(x))`$ for some generating $`J`$ Poisson pair $`(c_1,c_2)`$ (these dimensions do not depend on this pair, see Theorem 2). If this vector is independent of $`x`$ we call it a type of $`J`$ and say that $`J`$ is regular (cf. Example 3, below). If, moreover, all $`n_j,j=1,\mathrm{},k`$, are different we call $`J`$ simple. ###### 3.2. Definition Consider a manifold $`U`$ diffeomorphic to an open set in $`R^N`$, where $`N=(n_1+1)+\mathrm{}+(n_k+1),n_1<\mathrm{}<n_k`$, and a family $`𝒲=\{𝒲_\lambda \}_{\lambda P(𝒮)}`$ of $`k`$-codimensional foliations on $`U`$ parametrized by the projectivizaton of a two-dimensional vector space $`𝒮`$. We call $`𝒲`$ a Veronese web of type $`(n_1,\mathrm{},n_k)`$ if the following conditions are satisfied: there is a bundle filtration $`0=F_0\mathrm{}F_{k1}F_k=T^{}U`$ such that $`rankF_j/F_{j1}=n_j+1,j=1,\mathrm{},k`$; it induces the bundle filtration $`0=F_0𝒲_\lambda ^{}\mathrm{}F_{k1}𝒲_\lambda ^{}F_k𝒲_\lambda ^{}=𝒲_\lambda ^{},F_j𝒲_\lambda ^{}=F_j𝒲_\lambda ^{},j=1,\mathrm{},k`$, of the annihilating bundle $`𝒲_\lambda ^{}:=(T𝒲_\lambda )^{}T^{}U`$ so that $`rankF_j𝒲_\lambda ^{}=j`$; in particular $`A_j^\lambda (x):=F_j𝒲_{\lambda ,x}^{}/F_{j1}𝒲_{\lambda ,x}^{}`$ can be considered as a 1-dimensional subspace in $`A_j(x):=F_{j,x}/F_{j1,x}`$ for any $`xU`$; the map $`P(𝒮)\lambda A_j^\lambda (x)P(A_j(x))`$ is a Veronese inclusion for any $`xU,j=1,\mathrm{},k`$. ###### 3.3. Theorem Let $`J`$ be a simple bihamiltonian structure of type $`𝐧=(n_1,\mathrm{},n_k),n_1<\mathrm{}<n_k`$, and let $`xM`$ be a point of completeness for $`J`$. Write $`𝒱_\lambda `$ for the foliation of symplectic leaves of $`c_\lambda J`$. Then there exists a neighbourhood $`\stackrel{~}{U}x`$ such that $`U=\stackrel{~}{U}/`$ (see 1.9. Definition) is diffeommorphic to an open set in $`R^N`$ and $`\{𝒱_\lambda |_{\stackrel{~}{U}}/\}_{\lambda P(𝒮)}`$ is a Veronese web of type $`𝐧`$ on $`U`$. Proof. The theorem follows from Proposition 2.6. Definition. ###### 3.4. Example Let $`U=R^3(x,y,z)`$, $`\alpha _1=xdydz,\alpha _2^\lambda =\lambda _1dx+\lambda _2dy,k=2,n_1=0,n_2=1`$. Put $`\mathrm{\Gamma }(F_1)=Span\{\alpha _1\},\mathrm{\Gamma }((T𝒲_\lambda )^{})=Span\{\alpha _1,\alpha _2^\lambda \}`$, where $`\mathrm{\Gamma }`$ stands for the space of sections and $`Span`$ is taken over the ring of functions. Then $`\mathrm{\Gamma }(F_1(T𝒲_\lambda )^{})=\mathrm{\Gamma }(F_1)`$. Since $`T𝒲_\lambda TU`$ is a subbundle of rank 1, it is indeed tangent to 1-dimensional foliation $`𝒲_\lambda `$. Explicitely, $`\mathrm{\Gamma }(T𝒲_\lambda )=Span\{\lambda _2v_1\lambda _1v_2\}`$, where $`v_1=\frac{}{x},v_2=\frac{}{y}+x\frac{}{z}`$. On $`\stackrel{~}{U}=R(p)\times U`$ one defines the corresponding bihamiltonian structure as $`\{\frac{}{p}(\lambda _1v_1+\lambda _2v_2)\}_{(\lambda _1,\lambda _2)R^2}`$. ###### 3.5. Example Let $`U=R^5(x,y,z,s,t)`$, $`\alpha _1^\lambda =\lambda _1(xdydz)+\lambda _2(sdydt),\alpha _2^\lambda =\lambda _1^2dx+\lambda _1\lambda _2ds+\lambda _2^2dy,\mathrm{\Gamma }(F_1)=Span\{xdyds,sdydt\},\mathrm{\Gamma }((T𝒲_\lambda )^{}))=Span\{\alpha _1^\lambda ,\alpha _2^\lambda \},k=2,n_1=1,n_2=2`$. Then $`\mathrm{\Gamma }(F_1(T𝒲_\lambda )^{}))=Span\{\alpha _1^\lambda \}`$. The 3-distribution in $`TU`$ annihilating by 1-forms $`\alpha _1^\lambda ,\alpha _2^\lambda `$ is integrable since $$d\alpha _1^\lambda =\{\begin{array}{cc}\alpha _2^\lambda \frac{1}{\lambda _1}dy\hfill & \text{if}\lambda _10\hfill \\ \alpha _2^\lambda \frac{1}{\lambda _2}ds\hfill & \text{if}\lambda _1=0.\hfill \end{array}$$ Explicitely, $`\mathrm{\Gamma }(T𝒲_\lambda )=Span\{\lambda _2\frac{}{x}\lambda _1\frac{}{s},\lambda _2\frac{}{s}\lambda _1v,\lambda _2\frac{}{z}\lambda _1\frac{}{t}\}`$, where $`v=\frac{}{y}+x\frac{}{z}+s\frac{}{t}`$, and on $`\stackrel{~}{U}=R^3(p_1,p_2,p_3)\times U`$ the corresponding bihamiltonian structure is $`\{\frac{}{p_1}(\lambda _2\frac{}{x}\lambda _1\frac{}{s})+\frac{}{p_2}(\lambda _2\frac{}{s}\lambda _1v)+\frac{}{p_3}(\lambda _2\frac{}{z}\lambda _1\frac{}{t})\}_{(\lambda _1,\lambda _2)R^2}`$. ###### 3.6. Remark Off course, it is more convenient to describe Veronese webs in terms of a bundle direct decomposition $`B_1\mathrm{}B_k=T^{}U`$ such that $`F_j=_{i=1}^jB_i`$ rather than in terms of the isotypic filtration itself. In the above examples we used implicitely such a decomposition. However, one should remember that it is not unique. For instance, in Example 3.4. Example one has $`B_1=F_1,\mathrm{\Gamma }(B_2)=Span\{dx,ds,dy\}`$. But one could take $`\stackrel{~}{\alpha }_2^\lambda =\lambda _1^2(dx+xdydz)+\lambda _1\lambda _2(ds+sdydt)+\lambda _2^2dy=\alpha _2^\lambda +\lambda _1\alpha _1^\lambda `$ instead of $`\alpha _2^\lambda `$ and $`\mathrm{\Gamma }(\stackrel{~}{B}_2)=Span\{dx+xdydz,ds+sdydt,dy\}`$. Although this does not change the web, the corresponding decomposition is changed. In the author gives an involved analysis of this nonuniqueness. ###### 3.7. Definition A Veronese web admits the following local description. One can choose linear coordinates $`(\lambda _1,\lambda _2)`$ on $`𝒮`$ and a local coframe $`\alpha _1^1,\mathrm{},\alpha _{n_1+1}^1,\mathrm{},\alpha _1^k,\mathrm{},\alpha _{n_k+1}^k`$ such that $`\alpha _1^j,\mathrm{},\alpha _{n_j+1}^j\mathrm{\Gamma }(F_j),j=1,\mathrm{},k`$, and the annihilator $`(T𝒲_\lambda )^{}T^{}U`$ is generated by $`\alpha _\lambda ^1,\mathrm{},\alpha _\lambda ^k`$, where $`\alpha _\lambda ^j=\lambda _1^{n_j}\alpha _1^j+\lambda _1^{n_j1}\lambda _2\alpha _2^j+\mathrm{}+\lambda _2^{n_j}\alpha _{n_k+1}^j`$ (Veronese curve). If in a neighbourhood of any $`xU`$ there exists a holonomic coframe with the above properties, the Veronese web is called flat. In particular, all bundles in the isotypic filtration of a flat web are completely integrable as differential systems and, moreover, such a web splits to a direct product of flat Veronese webs of codimension $`1`$. The webs from Examples 3, 3.4. Example are not flat, since the bundles $`F_1`$ are nonintegrable. We conclude the section by an example of a complete bihamiltonian structure that is not regular. ###### 3.8. Example Let $`M=R^6`$ with coordinates $`(p_1,p_2,q_1,\mathrm{},q_4)`$, $`c_1=\frac{}{p_1}\frac{}{q_1}+\frac{}{p_2}\frac{}{q_2},c_2=\frac{}{p_1}(\frac{}{q_2}+q_1\frac{}{q_3})+\frac{}{p_2}\frac{}{q_4}`$. Here we have: two $`3`$-dimensional Kronecker blocks on $`MH,H=\{q_1=0\}`$; the $`5`$-dimensional Kronecker block and the $`1`$-dimensional zero block on the hyperplane $`H`$. ## 4 Veronese webs for the argument translation method. The notations from Subsection 1.11. Definition will be used below. We consider normal (déployable in terminology of Bourbaki, , IX,3) real form $`g`$ of complex simple Lie algebra. Let $`m_1,\mathrm{},m_r,r=rank(g)`$ be the exponents of $`g`$. ###### 4.1. Theorem Let $`(c_1,c_2)`$ be the Poisson pair from Example 1.11. Definition. Then the Veronese web $`\{𝒲_\lambda \}_{\lambda R^2}`$ of the corresponding bihamiltonian structure $`J`$ is of type $`(m_1,\mathrm{},m_r)`$ and is flat (Definition 3.6. Remark) in a neighbourhood of any point $`\pi (x)`$, where $`x(g^{}I)`$ and $`\pi `$ denotes the canonical projection $`\pi :g^{}I(g^{}I)/`$ (cf. 1.9. Definition, 3.2. Definition). Proof. Let $`g_1(x),\mathrm{},g_r(x),\mathrm{deg}g_j=m_j+1`$, be a set of algebraically independent global homogeneous polynomial Casimir functions for $`c`$ (see , VIII,8). Here we have identified $`g`$ and $`g^{}`$ by means of the Killing form. Note that $`g_1,\mathrm{},g_r`$ are functionally independent on $`gSingg`$, where $`Singg`$ is the set of adjoint orbits of nonmaximal dimension. Indeed, their restrictions to a Cartan subalgebra $`hg`$ are algebraically independent and invariant with respect to the Weyl group $`W`$. Now, we can apply the result of R.Steinberg () to deduce the nondegeneracy for the Jacobi matrix of $`g_1|_h,\mathrm{},g_r|_h`$ at a regular point. Consider the subspace $`d_0\mathrm{\Gamma }(T^{}g^{})`$ generated by the differentials of functions from the involutive set $`_0`$ (see 1) corresponding to $`J`$. It turns out that $`d_0`$ is generated by $`\{dg_j|_{\lambda _1x+\lambda _2a},(\lambda _1,\lambda _2)R^2,j=1,\mathrm{},r\}`$. If $`g_j^i(a,x),i=0,\mathrm{},m_j+1,j=1,\mathrm{},r`$, are the coefficients of the Taylor expansions $`g_j(x+\lambda a),j=1,\mathrm{},r`$, with respect to $`\lambda R`$, then one also has $$d_0=Span\{dg_j^i(a,x),i=0,\mathrm{},m_j,j=1,\mathrm{},r\}.$$ (4.1.1) Moreover, these differentials are linearly independent at any $`xg^{}I`$. This follows from the fact that $`J`$ is complete at $`g^{}I`$, from (4.1. Theorem), and from the formula $`_{j=1}^rm_j=\frac{1}{2}(dimgr)`$ (cf. , formula (F1), p. 289). Thus, we can regard $`g_j^i(a,x),i=0,\mathrm{},m_j,j=1,\mathrm{},r`$ as coordinates on the reduced space $`(g^{}I)/`$. Finally, $`(T𝒲_\lambda )^{},\lambda =(\lambda _1,\lambda _2)`$, is generated by $$\lambda _1^{m_j}dg_j^0(a,x)+\lambda _1^{m_j1}\lambda _2dg_j^1(a,x)+\mathrm{}+\lambda _2^{m_j}dg_j^{m_j}(a,x),j=1,\mathrm{},r.$$
warning/0001/gr-qc0001094.html
ar5iv
text
# 1 The bold curve corresponds to the noise power spectrum of VIRGO given in eq. (). The thin, dashed lines represent the minimal sensitivity required for the detection of the background at the 90% confidence level (i.e., 𝑆⁢𝑁⁢𝑅≃2.5), according to eq. (). The mass window compatible with detection corresponds to the range of frequency for which 𝑆̃ is below a given dashed line. ## Abstract We estimate the signal-to-noise ratio for two gravitational detectors interacting with a stochastic background of massive scalar waves. We find that the present experimental level of sensitivity could be already enough to detect a signal from a light but non-relativistic component of dark matter, even if the coupling is weak enough to exclude observable deviations from standard gravitational interactions, provided the mass is not too far from the sensitivity and overlapping band of the two detectors. BA-TH/00-374 January 2000 gr-qc/0001094 Signal-to-noise ratio for a stochastic background of massive relic particles M. Gasperini Dipartimento di Fisica, Università di Bari, Via G. Amendola 173, 70126 Bari, Italy and Istituto Nazionale di Fisica Nucleare, Sezione di Bari, Bari, Italy The sensitivity of present detectors to a stochastic background of relic gravitational waves has been recently discussed in detail in many papers (see -, for instance, and references therein). The sensitivity analysis has also been extended to include scalar waves , and scalar stochastic backgrounds of massless (or massive, but light enough) scalar particles, interacting with gravitational strength with the detectors. At present, however, no analysis seems to be available on the possible response of the gravitational antennas to a scalar stochastic background of non-relativistic particles. The aim of this paper is to compute the signal-to-noise ratio (SNR) for a pair of gravitational antennas by taking into account the possible mass of the background particles, in order to discuss in some detail the possible effects of the non-relativistic branch of their spectrum. We shall consider a cosmic stochastic background of massive scalar waves, whose energy density is coupled to the total mass of the detector with gravitational strength (or weaker). We shall assume that the background is characterized by a spectral energy density $`\mathrm{\Omega }(p)=d(\rho /\rho _c)/d\mathrm{ln}p`$, which we measure in units of critical density $`\rho _c=3H_0^2M_p^2/8\pi `$, and which extends in momentum space from $`p=0`$ to $`p=p_1`$ ($`p_1`$ is a cut-off scale depending on the details of the production mechanism). As a function of the frequency $`f=E(p)=(m^2+p^2)^{1/2}`$, the spectrum $`\stackrel{~}{\mathrm{\Omega }}(f)`$, $$\stackrel{~}{\mathrm{\Omega }}(f)\frac{d(\rho /\rho _c)}{d\mathrm{ln}f}=\left(\frac{f}{p}\right)^2\mathrm{\Omega }(p)$$ (1) thus extends over frequencies $`fm`$, from $`f=m`$ to $`f=f_1=(m^2+p_1^2)^{1/2}`$ (note that we are using “unconventional” units in which $`h=1`$, for a better comparison with the observable quantities used in the experimental analysis of gravitational antennas). We may thus distinguish three phenomenological possibilities. * $`mf_0`$, where $`f_0`$ is any frequency in the sensitivity band of the detector (tipically, if we are considering resonant masses and interferometers, $`f_010^210^3`$ Hz). In this case we expect no signal, as the response to the background should be totally suppressed by the intrinsic noise of the detector. * $`mf_0`$. In this case the detector, in its sensitivity band, responds to a relativistic frequency spectrum, and the SNR can be easily estimated by using the standard results. For a relativistic background of cosmological origin, however, the maximal amplitude allowed by nucleosynthesis is $`\mathrm{\Omega }10^5`$, possibly suppressed by a factor $`q^21`$ (in the interaction with the antenna) to avoid scalar-induced, long-range violations of the equivalence principle (see , for instance). We thus expect from such a scalar background a response not larger than from a background of relic gravitons, and then too weak for the sensitivity of present detectors. * $`mf_0`$. In this case the mass is the frequency band of maximal sensitivity, and the detector can respond resonantly also to the non-relativistic part of the background (i.e. to the branch $`p<m`$ of $`\mathrm{\Omega }(p)`$). In the non-relativistic sector, on the other hand, the background amplitude is not constrained by the nucleosynthesis bound, because the non-relativistic energy density grows in time with respect to the relativistic one: it could be sub-dominant at the nucleosynthesis epoch, even if today has reached a near-to-critical amplitude $`\mathrm{\Omega }1`$ (i.e., even if the massive background we are considering represents today a significant fraction of the cosmological dark matter). In such case, it will be shown in this paper that the present sensitivity of the existing gravitational antennas could be enough to distinguish the physical signal from the intrinsic experimental noise. We will follow the standard approach (see , for instance) in which the outputs of two detectors, $`s_i(t)`$, $`i=1,2`$, are correlated over an integration time $`T`$, to define a signal: $$S=_{T/2}^{T/2}𝑑t𝑑t^{}s_1(t)s_2(t^{})Q(tt^{}).$$ (2) Here $`Q(t)`$ is a real “filter” function, determined so as to optimize the signal-to-noise ratio (SNR), defined by an ensemble average as: $$SNR=S/\mathrm{\Delta }SS\left(S^2S^2\right)^{1/2}$$ (3) The outputs $`s_i(t)=h_i(t)+n_i(t)`$ contain the physical strain induced by the cosmic background, $`h_i`$, and the intrinsic instrumental noise, $`n_i`$. The two noises are supposed to be uncorrelated (i.e., statistically independent), $`n_1(t)n_2(t^{})=0`$, and much larger in magnitude than the physical strains $`h_i`$. Also, the cosmic background is assumed to be isotropic, stationary and Gaussian, with $`h_i=0`$. It follows that: $$S=_{T/2}^{T/2}𝑑t𝑑t^{}h_1(t)h_2(t^{})Q(tt^{}).$$ (4) An explicit compuation of the strain, at this point, would require a specific model of the interaction between the scalar background and the detector. We will assume in this paper that the strain $`h_i(t)`$, like in the case of gravitational waves and Brans-Dicke scalars , varies in time like the scalar fluctuation $`\varphi (x_i,t)`$ perturbing the detector (computed at the detector position $`x=x_i`$), and is proportional to the so-called “pattern function” $`F_i(\widehat{n})=e_{ab}(\widehat{n})D_i^{ab}`$, where $`\widehat{n}`$ is a unit vector specifying a direction on the two sphere, $`e_{ab}(\widehat{n})`$ is the polarization tensor of the scalar along $`\widehat{n}`$, and $`D_i^{ab}`$ is the detector tensor, specifying the orientation of the arms of the i-th detector. The field $`\varphi (x,t)`$ may represent the scalar (i.e, zero helicity) component of the metric fluctuations generated by the scalar component of the background, as in , or could even represent the background field itself, directly coupled to the detector through a “scalar charge” $`q_i`$ (for instance, a dilatonic charge), as discussed in . To take into account this second possibility, we shall explicitly introduce the scalar charge in the strain, by setting $$h_i(t)=q_i\varphi (x_i,t)e_{ab}(\widehat{n})D_i^{ab},$$ (5) where $`q_i=1`$ for scalar metric fluctuations, and $`q_i<1`$ for long-range scalar fields, phenomenologically constrained by the gravitational tests. The dimensionless parameter $`q_i`$ represents the net scalar charge per unit of gravitational mass of the detector, and is in general composition-dependent . To compute the average signal (4) we now expand the strain in momentum space, $`h_i(t)=q_i{\displaystyle 𝑑pd^2\widehat{n}\varphi (p,\widehat{n})F_i(\widehat{n})e^{2\pi i\left[p\widehat{n}\stackrel{}{x}_iE(p)t\right]}},`$ (6) $`p=|\stackrel{}{p}|,\stackrel{}{p}/p=\widehat{n},E(p)=f=(m^2+p^2)^{1/2},`$ (7) ($`d^2\widehat{n}`$ denotes the angular integral over the unit two-sphere), and we use the stochastic condition $$\varphi ^{}(p,\widehat{n}),\varphi (p^{},\widehat{n}^{})=\delta (pp^{})\delta ^2(\widehat{n}\widehat{n}^{})\mathrm{\Phi }(p).$$ (8) The isotropic function $`\mathrm{\Phi }(p)`$ can be expressed in terms of the spectral energy density $`\mathrm{\Omega }(p)`$, defined by $$\rho =\rho _cd\mathrm{ln}p\mathrm{\Omega }(p)=\frac{M_P^2}{16\pi }|\dot{\varphi }|^2,$$ (9) ($`M_P`$ is the Planck mass) from which: $$\mathrm{\Phi }(p)=\frac{3H_0^2\mathrm{\Omega }(p)}{8\pi ^3pE^2(p)}.$$ (10) By inserting the momentum expansion into eq. (4), and assuming, as usual, that the observation time $`T`$ is much larger than the typical time intervals $`tt^{}`$ for which $`Q0`$, we finally obtain: $$S=q_1q_2T\frac{2H_0^2}{5\pi ^2}\frac{dp}{pE^2(p)}\gamma (p)Q(p)\mathrm{\Omega }(p).$$ (11) We have defined the overlap function $`\gamma (p)`$ and the filter function $`Q(p)`$, in momentum space, as follows: $`\gamma (p)={\displaystyle \frac{15}{16\pi }}{\displaystyle d^2\widehat{n}F_1(\widehat{n})F_2(\widehat{n})e^{2\pi ip\widehat{n}(\stackrel{}{x}_2\stackrel{}{x}_1)}},`$ (12) $`Q(p)={\displaystyle 𝑑t^{}Q(tt^{})e^{2\pi iE(p)(tt^{})}}.`$ (13) Note that the overlap function depends on the relative distance of the two gravitational antennas and on their particular geometric configuration. In the above equation, in particular, $`\gamma (p)`$ has been normalized to the response of an interferometric detector to a scalar wave . We need now to compute the variance $`\mathrm{\Delta }S^2`$ which, for uncorrelated noises, much larger than the physical strains, can be expressed as : $$\mathrm{\Delta }S^2S^2=_{T/2}^{T/2}𝑑t𝑑t^{}𝑑\tau 𝑑\tau ^{}n_1(t)n_1(\tau )n_2(t^{})n_2(\tau ^{})Q(tt^{})Q(\tau \tau ^{}).$$ (14) It is convenient, in this context, to introduce the noise power spectrum in momentum space, $`S_i(p)`$, defined by $$n_i(t)n_i(\tau )=\frac{1}{2}𝑑pS_i(p)e^{2\pi iE(p)(t\tau )}.$$ (15) Assuming, as before, that $`T`$ is much larger than the typical correlation intervals $`tt^{}`$, $`\tau \tau ^{}`$, and using eq. (13) for $`Q(p)`$, then yields $$\mathrm{\Delta }S^2=\frac{T}{4}\frac{dp}{p}E(p)S_1(p)S_2(p)Q^2(p).$$ (16) The optimal filtering is now determined by the choice (see for details) $$Q(p)=\lambda \frac{\gamma (p)\mathrm{\Omega }(p)}{E^3(p)S_1(p)S_2(p)},$$ (17) where $`\lambda `$ is an arbitrary normalization constant. With such a choice we finally arrive, from eq. (11) and (17), to the optimized signal-to-noise ratio: $$SNR=\frac{S}{\mathrm{\Delta }S}=q_1q_2\frac{4H_0^2}{5\pi ^2}\left[T\frac{dp}{pE^5(p)}\frac{\gamma ^2(p)\mathrm{\Omega }^2(p)}{S_1(p)S_2(p)}\right]^{1/2}.$$ (18) It must be noted, at this point, that the functions $`S_i(p)`$ and $`\gamma (p)`$ appearing in the above equation are different, for a massive background, from the usual noise power spectrum $`\stackrel{~}{S}_i(f)`$, and overlap function $`\stackrel{~}{\gamma }(f)`$, conventionally used in the experimental analysis of gravitational antennas. Indeed, $`\stackrel{~}{S},\stackrel{~}{\gamma }`$ are defined as Fourier transforms of the frequency $`f=E(p)`$, so that (see for instance eq. (15)): $`{\displaystyle 𝑑f\stackrel{~}{S}_i(f)e^{2\pi ift}}={\displaystyle 𝑑pS_i(p)e^{2\pi iE(p)t}},`$ (19) $`{\displaystyle 𝑑f\stackrel{~}{\gamma }(f)e^{2\pi ift}}={\displaystyle 𝑑p\gamma (p)e^{2\pi iE(p)t}},`$ (20) from which $$S_i(p)=(df/dp)\stackrel{~}{S}_i(f),\gamma (p)=(df/dp)\stackrel{~}{\gamma }(f).$$ (21) By introducing into eq. (18) the known, experimentally meaningful variables $`\stackrel{~}{S}_i,\stackrel{~}{\gamma }`$, and using $`f=E(p)=(m^2+p^2)^{1/2}`$, we thus arrive at the final expression: $$SNR=q_1q_2\frac{4H_0^2}{5\pi ^2}\left[T_0^{p_1}\frac{d\mathrm{ln}p}{(m^2+p^2)^{5/2}}\frac{\mathrm{\Omega }^2(p)\stackrel{~}{\gamma }^2(\sqrt{m^2+p^2})}{\stackrel{~}{S}_1(\sqrt{m^2+p^2})\stackrel{~}{S}_2(\sqrt{m^2+p^2})}\right]^{1/2}.$$ (22) This equation represents the main result of this paper. For any given massive spectrum $`\mathrm{\Omega }(p)`$, and any pair of detectors with noise $`\stackrel{~}{S}_i`$ and overlap $`\stackrel{~}{\gamma }`$, the above equation determines the range of masses possibly compatible with a detectable signal ($`SNR>1`$), as a function of their coupling $`q_i`$ to the detectors. For $`m=0`$ we have $`p=f`$, and we recover the standard relativistic result , modulo a different normalization of the overlap function. For $`m0`$ we shall assume, as discussed at the beginning of this paper, that the mass lies within the sensitivity and overlapping band of the two detectors, i.e. $`\stackrel{~}{\gamma }(m)0`$, and $`\stackrel{~}{S}_i(m)`$ is near the experimental minimum. Also, let us assume that the non-relativistic branch of the spectrum, $`0<p<m`$, is near to saturate the critical density bound $`\mathrm{\Omega }<1`$, and thus dominates the total energy density of the background (the contribution of the relativistic branch $`p>m`$, if present, is assumed to be negligible). To estimate the integral of eq. (22), in such case, we can thus integrate over the non-relativistic modes only. In that range, we will approximate $`\stackrel{~}{S}_i`$ and $`\stackrel{~}{\gamma }`$ with their constant values at $`f=m`$. Assuming that the spectrum $`\mathrm{\Omega }(p)`$ avoids infrared divergences at $`p0`$ (like, for instance, a blue-tilted spectrum $`\mathrm{\Omega }(p)(p/p_1)^\delta `$, with $`\delta >0`$), we define $$_0^md\mathrm{ln}p\mathrm{\Omega }^2(p)=\mathrm{\Omega }_x^2,$$ (23) where $`\mathrm{\Omega }_x1`$ is a constant, possibly not very far from unity, and we finally arrive at the estimate $$SNRq_1q_2\frac{4H_0^2\mathrm{\Omega }_x}{5\pi ^2}\left[\frac{T\stackrel{~}{\gamma }^2(m)}{m^5\stackrel{~}{S}_1(m)\stackrel{~}{S}_2(m)}\right]^{1/2}.$$ (24) Following , the background can be detected, with a detection rate $`\gamma `$, and a false alarm rate $`\alpha `$, if $$SNR\sqrt{2}\left(\mathrm{erfc}^12\alpha \mathrm{erfc}^12\gamma \right).$$ (25) For a first qualitative indication, let us consider the ideal case in which the two detectors are coincident and coaligned, i.e. $`\stackrel{~}{\gamma }=1`$, $`\stackrel{~}{S}_1=\stackrel{~}{S}_2=\stackrel{~}{S}`$, $`q_1=q_2=q`$, and the massive stochastic background represents a dominant component of dark matter, i.e. $`\mathrm{\Omega }_xh_{100}^21`$ (where $`h_{100}=H_0/(100\mathrm{km}\mathrm{sec}^1\mathrm{Mpc}^1)`$ reflects the usual uncertainty in the present value of the Hubble parameter $`H_0`$). In such a case eq. (24), for an observation time $`T=10^8`$ sec, a detection rate $`\gamma =95\%`$, a false alarm rate $`\alpha =10\%`$, gives the condition: $$m^{5/2}\stackrel{~}{S}(m)<\frac{q^2}{3\pi ^2}10^{31}\mathrm{Hz}^{3/2}.$$ (26) We will use here, for a particular explicit example, the analytical fit of the noise power spectrum of VIRGO, which in the range from $`1`$ Hz to $`10`$ kHz can be parametrized as : $`\stackrel{~}{S}(f)`$ $`=`$ $`10^{44}\mathrm{sec}[3.46\times 10^6\left({\displaystyle \frac{f}{500\mathrm{Hz}}}\right)^5+6.60\times 10^2\left({\displaystyle \frac{f}{500\mathrm{Hz}}}\right)^1`$ (27) $`+`$ $`3.24\times 10^2+3.24\times 10^2\left({\displaystyle \frac{f}{500\mathrm{Hz}}}\right)^2].`$ (28) The intersection of this spectrum with the condition (26), in the plane $`\{\mathrm{log}\stackrel{~}{S},\mathrm{log}m\}`$, is shown in Fig. 1 for three possible values of $`q^2`$. The allowed mass window compatible with detection is strongly dependent on $`q^2`$, and closes completely for $`q^2<10^7`$, at least at the level of the noise spectrum used for this example. We should then consider two possibilities. If the spectrum $`\mathrm{\Omega }(p)`$ of eq. (22) refers to the spectrum of scalar metric fluctuations, induced on very small sub-horizon scales by an inhomogeneous, stochastic background of dark matter, then $`q^2=1`$ (since the detectors are geodesically coupled to metric fluctuations). In that case the detectable mass window extends over the full band from $`1`$ Hz to $`10`$ kHz, i.e from $`10^{15}`$ to $`10^{11}`$ eV. If, on the contrary, scalar metric fluctuations are negligible on such small scales, and $`\mathrm{\Omega }(p)`$ refers to the spectrum of the scalar background field itself, directly coupled to the detector through the scalar charge $`q`$, then this coupling is strongly suppressed in the mass range of Fig. 1, which corresponds to scalar interactions in the range of distance from $`10^6`$ to $`10^{10}`$ cm. Otherwise, such scalar field would induce long range corrections to the standard gravitational forces that would be detected in the precise tests of Newtonian gravity and of the equivalence principle (see for a complete compilation of the bounds on the coupling, as a function of the range). Taking into account all possible bounds , it follows that, if the scalar coupling is universal (i.e. the induced scalar force is composition-independent), then the maximal allowed charge $`q^2`$ is around $`10^7`$ from $`1`$ to $`10`$ Hz, and this upper bound grows proportionally to the mass (on a logarithmic scale) from $`10`$ to $`10^4`$ Hz. Composition-dependent couplings are instead more strongly constrained by Eotvos-like experiments: the maximal allowed value of $`q^2`$ scales like in the previous case, approximately, but the bounds are one order of magnitude stronger. By inserting into the condition (26) the gravitational bounds on $`q^2`$ we are led to the situation illustrated in Fig. 2. A scalar background of nearly critical density, non-universally coupled to macroscopic matter, turns out to be only marginally compatible with detection (at least, in the example illustrated in this paper), since the line of maximal $`q^2`$ is just on the wedge of the noise spectrum (28). If the coupling is instead universal (for instance, like in the dilaton model discussed in ), but the scalar is not exactly massless, then there is a mass window open to detection, from $`10^{14}`$ to $`10^{12}`$ eV. It seems appropriate to recall, at this point, that it is not impossible to produce a cosmic background of light, non-relativisic particles that saturates today the critical energy bound, as shown by explicit examples of spectra obtained in a string cosmology context . Such particles, typical of string cosmology, are in general very weakly coupled to the total mass of the detector (like the dilatons, if they are long range, and the charge of the antenna is composition-dependent), or even completely decoupled (like the axions, since the total axionic charge is zero for a macroscopic, unpolarized antenna). Nevertheless, it is important to stress that they could generate a spectrum of scalar metric fluctuations, gravitationally coupled to the detector, which follows the same non-relativistic behaviour of the original spectrum. We know, for instance, that in cosmological models based on the low-energy string effective action, the variable representing the dilaton fluctuations exactly coincides with the scalar part of the metric fluctuations (at least in an appropriate gauge ), and that the associated spectra also coincide. In view of the above discussion, the results illustrated in Fig. 1 and Fig. 2 suggest a new possible application of gravitational antennas, which seems to be interesting. Already at the present level of sensitivity, the gravitational detectors could be able to explore the possible presence of a light, massive component of dark matter, in a mass range that corresponds to their sensitivity band, in spite of the fact that such a massive background could be directly coupled to the total mass of the detector with a charge much weaker than gravitational, or only indirectly coupled, through the induced spectrum of scalar metric fluctuations. ###### Acknowledgements. It is a pleasure to thank Michele Maggiore and Gabriele Veneziano for interesting and useful discussions.
warning/0001/cond-mat0001223.html
ar5iv
text
# Self-Consistent Effective-Medium Approximations with Path Integrals ## I Introduction Among various effective-medium formulas used to model the effective behavior of random conducting linear composites, the symmetrical Bruggeman formula is undoubtly the most popular. Applied to an insulator/conductor binary mixture, it predicts a percolation-like transition for a volumic fraction of conductor $`p_c=1/d`$, where $`d`$ is the space dimension. The critical exponents are $`s=t=1`$, and its critical properties are thoroughly discussed in Ref. . This formula can be interpreted in two ways. On one hand, Milton has shown that it yields the exact effective conductivity of an ad hoc ideal medium built with a particular hierarchical structure. On the other hand, the Bruggeman formula can be seen as a first (one-body) self-consistent approximation to general disordered symmetric cell-materials , to which systematic corrections could be worked out. However, the Bruggeman approximation is very different from a mean-field theory of random conducting media. Indeed, an exact mean-field calculation on the Bethe lattice predicts a percolation threshold $`p_c1/(2d)`$ and exponents $`s=0`$ and $`t=3`$. These exponents are exact for $`d6`$, as well as the asymptotic behavior of the threshold when $`d\mathrm{}`$ (at least, for the hypercubic lattice to which a continuum theory naturally compares ). These values are also obtained in a more systematic mean-field theory for random resistor networks . The remarkable discrepancy between the mean-field results and Bruggeman’s formula indicates the ambiguous status of the Bruggeman theory. As a matter of fact, in spite of various (mostly perturbative) investigations in order to precise its theoretical status, the reasons for the peculiar critical behavior of Bruggeman’s formula are not completely cleared up. More surprisingly, another self-consistent effective-medium approximation due to Hori and Yonezawa (HY), obtained for the same type of media by means of a completely different approximation scheme (and later derived by functional methods ), exhibits the same exponents $`s=t=1`$ and a similar threshold behavior $`p_c=1\mathrm{exp}(1/d)1/d`$. Apart from phenomelogical variants, and up to our knowledge, the Bruggeman and HY effective-medium formulas are the only ones obtained from the equations of electrostatics in continuous media which are able to describe, at least qualitatively, the overall features of a percolation transition in any dimension. One intriguing question concerns the possibility of deriving alternative effective-medium formulas from a continuum formulation, which do not lead to the seemingly unavoidable values $`s=t=1`$ and $`p_c1/d`$. As we show in this paper, such a possibility exists. Our starting point is the path integral approach recently put forward by Barthélémy and Orland, where the effective-medium problem is recast in a functional form. The problem reduces to compute a free-energy: roughly, the logarithm, averaged over the disorder, of a functional integral of Boltzmann-like weights, over allowed field configurations (which include boundary conditions). The average of the logarithm is carried out with the replica method (already used in Ref. ). In Ref. , the authors showed that the path integral formulation allows one to easily recover the second-order weak-disorder expansion of the effective permittivity of nonlinear composites . However, this formulation has not yet been used to derive self-consistent estimates. In this paper, we show how this can be done. After a presentation of the functional approach to the homogeneization problem, and of the replica method (Sec. II), we discuss self-consistent effective-medium approximations (Sec. III). As usual in such approximations, a background reference medium is introduced under the form of an ansatz for the energy of the system, whose parameters are to be determined self-consistently (Sec. III A). The new feature here is that the ansatz contains a replica-coupling term, whose significance is explained (Sec. III B). The self-consistency conditions to determine its parameters are next discussed, and two types of effective-medium formulas are identified (Sec. III C): one in which the replica couplings are cancelled (hereafter referred to as “type 1”), and the other one with non-zero replica couplings (“type 2”). Two different approximations are then worked out for each type (Sec. IV). It is found that type 1 generates the Bruggeman and HY formulas, whereas type 2 brings in two new effective-medium formulas which are “replica coupling counterparts” of the previous ones. They possess exponents $`s=0`$, $`t=2`$, and a threshold $`p_c1/(2d)`$ (Sec. IV B 3). These new formulas are discussed in Sec. V, where numerical results are presented before we conclude in Sec. VI. ## II Path integral formulation of the problem The effective properties of a random conducting medium can be defined with the help of the total dissipated power in the medium. In terms of the electric field $`E(x)`$, the dissipated power $`w`$ in the system of volume $`V`$ reads $$W[E]=_V𝑑xw_x\left(E(x)\right),$$ (1) where $`w_x`$ is the local power density. Hereafter, we take the volume $`V`$ of the sample equal to one. In heterogeneous materials, $`w_x`$ depends on constitutive parameters randomly varying from point to point. For linear conducting media with $`j(x)=\sigma (x)E(x)`$, where $`\sigma `$ is the local random conductivity and $`j`$ is the electric current, we have $$w_x\left(E(x)\right)=\sigma (x)E^2(x)/2.$$ (2) In the analogous effective permittivity problem the dissipated power is replaced by the stored energy $`\epsilon (x)E^2(x)/2`$ ($`\epsilon `$ is the permittivity). For this reason, we shall abusively refer to $`w_x`$ as the “energy density” hereafter. In the nonlinear problem, $`w_x(E)`$ is a non-quadratic function of $`E`$. An alternative to solving Maxwell’s equations is to minimize the total energy $`W`$ subjected to the two constraints : (i) $`E=\varphi `$ and (ii) $`\overline{E}=E_0`$; here, the bar stands for a spatial average, and $`E_0`$ is a constant applied electric field. The minimum, $`W^{}\left(E_0\right)`$, is expected to be self-averaging, as occurs for the free-energy in disordered systems. We can therefore write $$W^{}(E_0)=\underset{\genfrac{}{}{0pt}{}{\overline{E}=E_0}{E=\varphi }}{\text{min}}W[E]$$ (3) where the brackets $``$ denote the disorder average. $`W^{}(E_0)`$ is the energy in a homogeneous medium characterized by an effective constitutive law $$j=\frac{W^{}(E_0)}{E_0}=\sigma _{\text{eff}}E_0.$$ (4) The second equality defines the effective conductivity of the medium. The problem thus reduces to computing the average of the constrained minimum of a functional of the electric field. The electric field derives from a potential and has a fixed mean value. We can rewrite the constrained minimum in (3) using a path integral $$\underset{\genfrac{}{}{0pt}{}{\overline{E}=E_0}{E=\varphi }}{\text{min}}W[E]=\underset{\beta \mathrm{}}{lim}\frac{1}{\beta }\text{ln}𝒟E𝒟\varphi \delta (E+\varphi )\delta (\overline{E}E_0)e^{\beta W[E]}.$$ (5) The minimum can be interpreted as the ground state energy associated to the partition function $$Z=\stackrel{~}{𝒟}Ee^{\beta W[E]},$$ (6) where we have used the shorthand notation $$\stackrel{~}{𝒟}E=𝒟E\delta (\overline{E}E_0)𝒟\varphi \delta (E+\varphi )$$ (7) for the constrained functional measure. We need to compute the average of the logarithm of (6). In order to proceed, we introduce replicas and use the identity $`\mathrm{ln}Z=lim_{n0}(Z^n1)/n`$, hence $$W^{}=\underset{\beta \mathrm{}}{lim}\underset{n0}{lim}\frac{1}{n\beta }(Z^n1).$$ (8) The limits do not commute. The equivalent form $$W^{}=\underset{\beta \mathrm{}}{lim}\underset{n0}{lim}\frac{1}{n\beta }\mathrm{ln}Z^n$$ (9) can be used as well. The replica method relies on the fact that one can easily compute the replicated partition function $`Z^n`$ for $`n`$ integer, and subsequently take the limit $`n0`$. The main quantity of interest therefore is $$Z^n=\underset{\alpha =1}{\overset{n}{}}\stackrel{~}{𝒟}E^\alpha e^{\beta _{\alpha =1}^nW[E^\alpha ]}.$$ (10) Denoting the replicated measure by $`\stackrel{~}{𝒟}\left(E^\alpha \right)=_{\alpha =1}^n\stackrel{~}{𝒟}E^\alpha `$, the average $`Z^n`$ can be written in terms of an “effective Hamiltonian” $$Z^n=\stackrel{~}{𝒟}\left(E^\alpha \right)e^{\beta _e},$$ (11) with $$_e=\frac{1}{\beta }\mathrm{ln}e^{\beta _{\alpha =1}^nW[E^\alpha ]}.$$ (12) For simplicity, we restrict ourselves to cell materials where the local properties are statistically uncorrelated from site to site. Volume integrals may then to be identified with sums over sites (each pertaining to one cell) according to the correspondence $`𝑑xv_x`$, where $`v`$ is an infinitesimal cell volume (which defines the microscopic correlation length of the problem). Then, $`_e`$ simplifies to $$_e=\frac{1}{\beta }\frac{dx}{v}\mathrm{ln}e^{\beta v_\alpha w_x\left(E^\alpha (x)\right)}.$$ (13) Note that our discussion in Sec. III will be specialized to binary disorder for which the constitutive parameters can take only two values (but the proofs are general). That is, we assume that the local energy density is distributed according to the probability distribution $$P\left(w=w_x(E)\right)=p\delta \left(ww_1(E)\right)+q\delta \left(ww_2(E)\right).$$ (14) (where $`q=1p`$). With this choice, $$_e=\frac{1}{\beta }\frac{dx}{v}\mathrm{ln}\left[pe^{\beta v_{\alpha =1}^nw_1\left(E^\alpha (x)\right)}+qe^{\beta v_{\alpha =1}^nw_2\left(E^\alpha (x)\right)}\right].$$ (15) The above formalism applies to any form of the energy density, and in particular to nonlinear media . A method for extracting from the path integral the second-order weak-contrast perturbation expansion of the effective potential $`W^{}(E_0)`$, for nonlinear media, has been introduced in Ref. . ## III Principle of self-consistent approximations In this paper, we consider the linear problem only. This section is devoted to self-consistent approximations to $`W^{}`$. We first present the principle for building such approximations through the introduction of a trial Hamiltonian. Then, we discuss the choice of a trial Hamiltonian with replica couplings. Finally, we explain how to exploit these replica couplings in order to obtain two kinds of self-consistent formulas. ### A Overview The common ingredient to the approximations discussed below is the introduction of a linear comparison medium described by a trial Hamiltonian $`_0`$ which is quadratic in the electric field and non-random, e.g. the one-parameter ansatz $$_0=\frac{\sigma _0}{2}𝑑x\underset{\alpha }{}E_{}^{\alpha }{}_{}{}^{2}(x),$$ (16) where $`\sigma _0>0`$ is to be determined by an appropriate self-consistency condition. This Hamiltonian is that of a (replicated) homogeneous medium, but without couplings between replicas. Its meaning and that of other possible choices with replica couplings are discussed below. The partition function $`Z^n`$ can be rewritten as $$Z^n=\frac{\stackrel{~}{𝒟}\left(E^\alpha \right)e^{\beta (_e_0)}e^{\beta _0}}{\stackrel{~}{𝒟}\left(E^\alpha \right)e^{\beta _0}}\stackrel{~}{𝒟}\left(E^\alpha \right)e^{\beta _0},$$ (17) or, with another notation $$Z^n=e^{\beta (_e_0)}_0Z_0,$$ (18) where $`Z_0`$ is the partition function associated to $`_0`$, and $`_0`$ stands for the functional average with weights $`e^{\beta _0}/Z_0`$. Equ. (9) thus reads $$W^{}=W_0+\mathrm{\Delta }W,$$ (19) where $`W_0(E_0)`$ $`=`$ $`\underset{\genfrac{}{}{0pt}{}{n0}{\beta \mathrm{}}}{lim}{\displaystyle \frac{1}{n\beta }}\mathrm{ln}Z_0,`$ (20) $`\mathrm{\Delta }W(E_0)`$ $`=`$ $`\underset{\genfrac{}{}{0pt}{}{n0}{\beta \mathrm{}}}{lim}{\displaystyle \frac{1}{n\beta }}\mathrm{ln}e^{\beta (_e_0)}_0.`$ (21) The quantity $`\mathrm{\Delta }W(E_0)`$ is difficult to compute (an exact evaluation would lead to the exact result for the effective conductivity), and we have to resort to approximations. A natural self-consistency condition for $`_0`$ is $$\mathrm{\Delta }W(E_0)=0,$$ (22) which completely determines $`_0`$ in the case where it depends on one single parameter, as in (16). For more general choices of $`_0`$ with several free parameters, (22) only provides a relation between these parameters, and additional considerations are in order to determine them all. First of all, we have to precise the form of the ansatz to be used in our calculations. ### B Replica couplings and choice of the ansatz We deduce here the form of the trial Hamiltonian $`_0`$ from an analysis of the effective Hamiltonian $`_e`$. Eq. (12) shows that the effective Hamiltonian is non-random but that the average over disorder introduced a coupling between different replicas. The meaning of these couplings is more transparent if we carry out an expansion of (12) around the average field $`\overline{E}=E_0`$ as in the weak-contrast expansion . With $`_i=/E_i`$ and $`\mathrm{\Delta }E^\alpha =E^\alpha E_0`$ we have $`_e`$ $`=`$ $`nw_x(E_0)`$ (23) $`+`$ $`{\displaystyle \frac{1}{2}}\left[{\displaystyle \underset{\alpha }{}}{\displaystyle 𝑑xa_{ij}\mathrm{\Delta }E_i^\alpha (x)\mathrm{\Delta }E_j^\alpha (x)}\beta {\displaystyle \underset{\alpha ,\gamma }{}}{\displaystyle 𝑑x𝑑yc_{ij}^{(2)}(xy)\mathrm{\Delta }E_i^\alpha (x)\mathrm{\Delta }E_j^\gamma (y)}\right]+\mathrm{},`$ (24) where $`a_{ij}`$ $`=`$ $`_{ij}^2w_x(E_0),`$ (25) $`c_{ij}^{(2)}(xy)`$ $`=`$ $`_iw_x(E_0)_jw_y(E_0)_iw_x(E_0)_jw_y(E_0).`$ (26) The first non-zero replica-coupling term is proportional to $`\beta c^{(2)}`$. We thus see that the coupling between replicas acts only within clusters defined by $`n`$-point connected correlation functions $`c^{(n)}`$, and accounts for the fluctuations of the electric field in these clusters. The replica coupling would vanish if there were no disorder at all. In the limit where the size of the region defined by $`c^{(2)}`$ shrinks to zero – which means that the system is observed at a macroscopic level, we can approximate $$c_{ij}^{(2)}(xy)vc_{ij}^{(2)}(0)\delta (xy),$$ (27) and we recover the expansion $$_e=nw_x(E_0)+\frac{1}{2}𝑑x\left[\underset{\alpha }{}a_{ij}\mathrm{\Delta }E_i^\alpha (x)\mathrm{\Delta }E_j^\alpha (x)v\beta \underset{\alpha ,\gamma }{}c_{ij}^{(2)}(0)\mathrm{\Delta }E_i^\alpha (x)\mathrm{\Delta }E_j^\gamma (x)\right]+\mathrm{}.$$ (28) which could directly be obtained from (13). The presence of $`v`$ in front of the replica coupling term is the macroscopic remnant of a microscopic average having been taken within a two-particle cluster, of center $`x`$ and volume $`v`$. This discussion therefore enlightens a relation between replica coupling and the electric field fluctuations within clusters. Expansion (28) suggests a two-parameter replica-symmetric ansatz of the form $$_0=\frac{1}{2}\underset{\alpha \gamma }{}𝑑xM^{\alpha \gamma }E_i^\alpha E_i^\gamma ,$$ (29) where $$M^{\alpha \gamma }=\sigma _0\delta _{\alpha \gamma }v\beta QE_0^2.$$ (30) The free parameters are $`\sigma _0`$ and $`Q`$. Note that $`Q`$ has the dimension of a squared conductivity, because it is related to a quantity relative to two points. For simplicity, the ansatz $`M`$ is diagonal in the euclidean vector space. However, we tried calculations with a tensorial structure reproducing that of $`a_{ij}`$ and $`c_{ij}^{(2)}`$ in Eqs. (25), (26); but, apart from a different normalization for $`Q`$, no differences showed up in the final effective-medium theories (as far as linear media are concerned). An interesting feature of the ansatz (30) is that, though being non-random, it embodies underlying disorder through its replica couplings. In order to understand this point, we compute $`W_0(E_0)`$ given by (20) $$W_0(E_0)=\underset{\genfrac{}{}{0pt}{}{n0}{\beta \mathrm{}}}{lim}\frac{1}{n\beta }\mathrm{ln}\stackrel{~}{𝒟}Ee^{\frac{\beta }{2}{\scriptscriptstyle 𝑑x\left(\sigma _0_\alpha E_{}^{\alpha }{}_{}{}^{2}v\beta QE_0^2_{\alpha ,\gamma }E^\alpha E^\gamma \right)}}.$$ (31) After writing $`E=E_0\varphi `$, and going to the Fourier transform of $`\varphi `$ , we arrive at $`W_0(E_0)`$ $`=`$ $`\underset{\genfrac{}{}{0pt}{}{n0}{\beta \mathrm{}}}{lim}{\displaystyle \frac{1}{n\beta }}\mathrm{ln}\left[(\text{Det}M)^{1/2v}e^{\frac{\beta }{2}_{\alpha \gamma }M_{\alpha \gamma }E_0^2}\right]`$ (32) $`=`$ $`{\displaystyle \frac{1}{2}}\sigma _0(1Q/\sigma _0^2)E_0^2.`$ (33) Carrying out the derivative of (31) with respect to $`\sigma _0`$, we obtain $$\underset{\genfrac{}{}{0pt}{}{n0}{\beta \mathrm{}}}{lim}\frac{1}{n}\underset{\alpha }{}E_{}^{\alpha }{}_{}{}^{2}_0=E_0^2+\frac{Q}{\sigma _0^2}E_0^2,$$ (34) where volume averages $`\overline{E^2}`$ have been replaced by statistical ones, the microscopic size $`v^{1/d}`$ being much smaller than that of the system, $`V^{1/d}=1`$. All the replicas are equivalent, and the functional average $`_0`$ selects in the limit $`\beta \mathrm{}`$ the real field in the medium. Hence, setting $`\mathrm{\Delta }E=EE_0`$, the previous equation leads to $$\frac{\mathrm{\Delta }E^2}{E_0^2}=\frac{Q}{\sigma _0^2}.$$ (35) which implies that $`Q0`$. When $`Q0`$, the electric field fluctuates in the medium, whereas it is uniform when $`Q=0`$. The ansatz $`_0`$ therefore represents a medium which is homogeneized (because it is non-random), but which nonetheless accounts for field fluctuations. We thus expect new effective medium approximations when the replica coupling $`Q`$ is non zero. ### C Self-consistency Up to this point the discussion focused on the ansatz itself, without referring to $`_e`$. In particular, $`\sigma _0`$ was treated as a mere number. We now discuss what happens when self-consistency is imposed, within some approximation scheme. The medium is made of $`N`$ phases labelled by $`\nu `$, of respective conductivities $`\sigma _\nu `$ and volume concentrations $`p_\nu `$. The self-consistency relation $`\mathrm{\Delta }W(E_0)=0`$, which imposes constraints on the ansatz, determines $`Q`$ as a function $`Q=Q(\sigma _0,\{\sigma _\nu \})`$. Then $`W^{}=W_0`$ and, with (4), $$\sigma _{\text{eff}}=\sigma _0\left[1\frac{Q(\sigma _0,\{\sigma _\nu \})}{\sigma _0^2}\right].$$ (36) Suppose now that $`\sigma _0=\sigma _0(\{\sigma _\nu \})`$ is determined by an additional condition (to be precised below). Using the exact formula (cf. Appendix A) $$\frac{\mathrm{\Delta }E^2}{E_0^2}=\underset{\nu }{}\frac{\sigma _{\text{eff}}}{\sigma _\nu }1,$$ (37) the fluctuations of the electric field deduced from (36) can be written $$\frac{\mathrm{\Delta }E^2}{E_0^2}=\sigma _{\text{eff}}^{}(\sigma _0)\left(\underset{\nu }{}\frac{\sigma _0}{\sigma _\nu }1\right)+\frac{Q}{\sigma _0^2}\frac{1}{\sigma _0}\left[\frac{Q}{\sigma _0}+\underset{\nu }{}\left(\frac{Q}{\sigma _\nu }\right)_{\sigma _0}\right]$$ (38) where the last derivative is performed at constant $`\sigma _0`$. This expression distinguishes between different contributions to the field fluctuations: (i) the first term represents fluctuations coming from the “macroscopic” background effective medium $`\sigma _0`$; (ii) the second one is that already found in (35), and would be the only one if $`Q`$ were independent from $`\sigma _0`$, and if $`\sigma _0`$ were equal to $`\sigma `$, the trivial value corresponding to a non-fluctuating reference medium for a multiphase composite, cf. Appendix A; (iii) and finally a third term comes from the dependence of $`Q`$ on $`\sigma _0`$ and $`\sigma _\nu `$. Both last terms are, according to the interpretation of replica coupling developped in the previous section, of “microscopic” origin. We now turn to the determination of $`\sigma _0(\{\sigma _\nu \})`$. A first obvious self-consistency condition for $`\sigma _0`$ is $`Q0`$, so that $`\sigma _{\text{eff}}=\sigma _0`$. The effective-medium formulas obtained this way are referred to as “type 1” hereafter. As is shown below, to this type pertain the Bruggeman and HY formulas. “Type 2” effective-medium formulas are obtained by taking $`\sigma _0`$ as the solution of $`\sigma _{\text{eff}}^{}(\sigma _0)=0`$, and by using this value in $`\sigma _{\text{eff}}`$. According to (38), this procedure makes the effective-medium insensitive to the fluctuations generated in the reference medium $`\sigma _0`$, so that relevant fluctuations only come from $`Q`$. ## IV Two approximations In this section, the ideas introduced above are used within two different approximations to $`\mathrm{\Delta }W(E_0)`$, based on the ansatz (29), (30). For each approximation to $`\mathrm{\Delta }W`$, “type 1” and “type 2” formulas are obtained. Herafter, $`q=Q/\sigma _0^2`$, so that $$\sigma _{\text{eff}}=\sigma _0(1q).$$ (39) ### A One-impurity approximation We first consider a “one-impurity” (or “local”) calculation. The Bruggeman formula emerges as the “type 1” effective-medium formula in this approximation, which is not suprising since it can be seen as a one-site (self-consistent) theory . #### 1 Approximation scheme The approximation for $`\mathrm{\Delta }W(E_0)`$ \[Eq. (21)\], detailed in Appendix B, is a one-impurity approximation where interactions between different points are ignored. Let us denote by $`w_0`$ the trial Hamiltonian density, which depends on all the replicas, defined from (29), (30) by $$_0𝑑xw_0[E(x)].$$ (40) Here and in Appendix B, the notation $`[]`$ indicates a dependence with respect to all the replicas. Setting $$\mathrm{\Delta }w_x[E(x)]=\underset{\alpha }{}w_x\left(E^\alpha (x)\right)w_0[E(x)],$$ (41) the one-impurity approximation results in $$e^{\beta (_e_0)}_01+\frac{1}{v}e^{\beta v\mathrm{\Delta }w_x[E(x)]}_01.$$ (42) Because of statistical translation invariance, the final result is independant of the point $`x`$. The right-hand side can be computed exactly for any potential $`w_x`$ in the limit $`\beta \mathrm{}`$ using a saddle-point method. Setting $`\mathrm{\Delta }\sigma =\sigma \sigma _0`$ and $$\mu =\left(1+\frac{\mathrm{\Delta }\sigma }{d\sigma _0}\right)^1,$$ (43) we arrive at $$\mathrm{\Delta }W(E_0)=\frac{1}{2}\mathrm{\Delta }\sigma \mu E_0^2+\frac{q}{2}\sigma \mu E_0^2.$$ (44) The condition $`\mathrm{\Delta }W(E_0)=0`$ yields $$q=\frac{\mathrm{\Delta }\sigma \mu }{\sigma \mu }.$$ (45) #### 2 Type 1 formula: Bruggeman’s Letting $`q0`$ amounts to imposing $`\mathrm{\Delta }\sigma \mu =0`$, which is nothing but the Bruggeman equation $$\frac{\sigma \sigma _0}{\sigma +(d1)\sigma _0}=0.$$ (46) The Bruggeman equation can also be written $`\mu =1`$, or $`\sigma _0=\sigma \mu `$ if $`d1`$, or $`\sigma _0=\sigma \mu /\mu `$. The last expression is suitable for computing $`\sigma _0`$ iteratively (starting, e.g., from $`\sigma _0=\sigma `$) in any dimension. The Bruggeman conductivity $`\sigma _{\text{eff}}=\sigma _0`$ possesses a percolation threshold $`p_c=1/d`$, and critical exponents $`s=t=1`$ . The fluctuations computed from (37) read $$\frac{\mathrm{\Delta }E^2}{E_0^2}=\frac{\sigma _0\mu ^2}{\sigma \mu ^2}.$$ (47) #### 3 Type 2 formula We now let $`q0`$ and given by (45) and $`\sigma _{\text{eff}}(\sigma _0)=\sigma _0\left(1q(\sigma _0)\right)`$. The equation $`\sigma _{\text{eff}}^{}(\sigma _0)=0`$ reads $$\sigma _0=\frac{\sigma \mu }{\mu }\left(1+\frac{\mu \sigma ^2\mu ^2\sigma \mu ^2\sigma \mu }{2d\sigma \mu ^2}\right).$$ (48) Like Bruggeman’s, this equation is easily solved by iterations starting from $`\sigma _0=\sigma `$. The iterations then always converge to the physical solution, which we denote by $`\sigma _0^{}`$. The effective conductivity thus is $`\sigma _{\text{eff}}=\sigma _0^{}\left(1q(\sigma _0^{})\right)`$. To study its critical behavior, we consider a binary mixture, where $`\sigma =\sigma _1`$ with probability $`(1p)`$, and $`\sigma =\sigma _2`$ with probability $`p`$. In the conductor/superconductor limit where $`\sigma _2\mathrm{}`$ we find, setting $`p_c=1/(2d1)`$, $$\sigma _0^{}=\frac{\sigma _1}{p(d1)}\left(\sqrt{\frac{1p}{1p/p_c}}1\right)(p<p_c)$$ (49) and $$\sigma _{\text{eff}}=2\sigma _1\frac{\left[1dp\sqrt{(1p)(1p/p_c)}\right]}{p^2(d1)^2}(p<p_c).$$ (50) The critical concentration $`p_c`$ can be interpreted as a percolation threshold, and is the same as that obtained in the mean-field model on a Bethe lattice with connectivity $`z=2d`$. Since $`\sigma _{\text{eff}}=2(2d1)\sigma _1/(d1)(p_cp)^0`$ for $`pp_c`$, the superconductivity exponent is $`s=0`$. Note however that $`\sigma _0`$ displays a square-root cusp at $`p=p_c`$. The critical behavior for $`p>p_c`$ is obtained by exmining the insulator/conductor mixture where $`\sigma _2`$ is finite and $`\sigma _1=0`$. Then $`\sigma _0^{}`$ $`=`$ $`\sigma _2{\displaystyle \frac{p/p_c1}{2(d1)}},`$ (51) $`\sigma _{\text{eff}}`$ $`=`$ $`\sigma _2{\displaystyle \frac{(p/p_c1)^2}{4(d1)^2p}}(p>p_c).`$ (52) Since $`\sigma _{\text{eff}}(pp_c)^2`$ for $`pp_c`$, the conductivity exponent is $`t=2`$. For the special case of $`d=1`$, $`\mu =\sigma _0/\sigma `$ so that (48) reduces to $`\sigma _0=1/\sigma ^1`$, and $`q=0`$. Therefore, $`\sigma _{\text{eff}}=1/\sigma ^1`$, which is the exact result. Like Bruggeman’s, the new formula is also exact to second order in the contrast, in any dimension $$\sigma _{\text{eff}}=\sigma \left[1\frac{\sigma ^2\sigma ^2}{d\sigma ^2}+\mathrm{}\right];$$ (53) and in the dilute limit where (e.g.) $`p_21`$ $$\sigma _{\text{eff}}=\sigma _1\left[1+d\frac{\sigma _2\sigma _1}{\sigma _2+(d1)\sigma _1}+\mathrm{}\right].$$ (54) In the discussion (Section V), it is argued that because of its exponents $`st`$, and because it is less trivial than the Bruggeman formula (especially in the insulator/conductor case where it does not reduce to a straight line), this formula may constitute an easy-to-handle alternative to the latter in dimensions $`d3`$. Graphical comparisons between different effective-medium formulas are discussed in Sec. V. ### B Cumulant series approximation In this section, we show how to recover by means of a cumulant approximation to $`\mathrm{\Delta }W`$ the effective-medium formula of HY, together with its “type 2” counterpart. #### 1 Approximation scheme We consider the first-order cumulant approximation $$e^{\beta (_e_0)}_0e^{\beta _e_0_0}.$$ (55) We have then $$\mathrm{\Delta }W(E_0)\underset{\genfrac{}{}{0pt}{}{n0}{\beta \mathrm{}}}{lim}\frac{1}{n}_e_0_0.$$ (56) As is shown in Appendix C, the calculations here involve an expansion in a series of the cumulants of the distribution of $`\sigma `$, whose significance has been discussed at length in the original paper by HY. After some algebra, we obtain (cf. Appendix C) $$\mathrm{\Delta }W(E_0)=\frac{\sigma _0}{2}\left\{\left[1+dh_0\left(1/(d\sigma _0)\right)\right]+dqh_0\left(1/(d\sigma _0)\right)\right\}E_0^2,$$ (57) where $$h_0(z)=_0^{\mathrm{}}𝑑ue^u\mathrm{ln}e^{u\sigma z}.$$ (58) The family of functions $`h_k`$ is defined in Appendix C. The self-consistency $`\mathrm{\Delta }W(E_0)=0`$ now yields $$q=\left[1+\frac{1}{dh_0\left(1/(d\sigma _0)\right)}\right].$$ (59) #### 2 Type 1 formula: the Hori-Yonezawa formula We first consider the case with no couplings between replicas, i.e. $`q=0`$. The HY formula for $`\sigma _0`$ reads $$h_0\left(1/(d\sigma _0)\right)=\frac{1}{d},$$ (60) and the effective conductivity is $`\sigma _{\text{eff}}=\sigma _0`$. It can be shown that $`\sigma _{\text{eff}}`$ displays a percolation threshold $`p_c=1\mathrm{exp}(1/d)`$, and exponents $`s=t=1`$. Applying (37) and using (C14), the fluctuations read $$\frac{\mathrm{\Delta }E^2}{E_0^2}=\frac{2+dh_1\left(1/(d\sigma _0)\right)}{1+dh_1\left(1/(d\sigma _0)\right)}.$$ (61) #### 3 Type 2 formula We now consider $`q0`$ and determined as a function of $`\sigma _0`$ by (59). Then $$\sigma _{\text{eff}}(\sigma _0)=\sigma _0\left[2+\frac{1}{dh_0\left(1/(d\sigma _0)\right)}\right].$$ (62) The equation for $`\sigma _0`$ is $`\sigma _{\text{eff}}^{}(\sigma _0)=0`$; that is, with (C14) $$2+\frac{1}{d}\frac{h_1\left(1/(d\sigma _0)\right)}{h_0^2\left(1/(d\sigma _0)\right)}=0.$$ (63) In order to study the critical behavior of $`\sigma _{\text{eff}}`$, we consider again a binary mixture where $`\sigma =\sigma _1`$ with probability $`(1p)`$, and $`\sigma =\sigma _2`$ with probability $`p`$. In the conductor/superconductor case where $`\sigma _2\mathrm{}`$, we have $`h_0(1/(d\sigma _0)))=\mathrm{ln}(1p)\sigma _1/(d\sigma _0)`$ and a similar equation for $`h_1`$, so that (63) reduces to a second-degree polynomial equation. Its physical solution reads $$\sigma _0^{}=\frac{2\sigma _1}{\sqrt{1+2d\mathrm{ln}(1p)}\left[1+\sqrt{1+2d\mathrm{ln}(1p)}\right]}.$$ (64) It is defined for $`p`$ less than a critical value $$p_c=1e^{1/(2d)}.$$ (65) This percolation threshold is the same as the one obtained in the Potts model at the mean-field level, and in the mean-field theory of Ref. . Reporting (64) into (62), we arrive at $$\sigma _{\text{eff}}=\frac{4\sigma _1}{\left[1+\sqrt{1+2d\mathrm{ln}(1p)}\right]^2}(p<p_c).$$ (66) Since $`\sigma _{\text{eff}}(p_cp)^0`$ for $`pp_c`$, the superconductivity exponent is $`s=0`$. In the opposite insulator/conductor case, where $`\sigma _1=0`$ and $`\sigma _2`$ is finite, the solution for $`p>p_c`$ can only be found perturbatively around the percolation threshold. Expanding the logarithm in $`h_0`$ and $`h_1`$ as $$\mathrm{ln}\left[(1p)+pe^{u\sigma _2/(d\sigma _0)}\right]=\mathrm{ln}(1p)+\underset{l1}{}\frac{(1)^{l1}}{l}\left(\frac{p}{1p}\right)^le^{lu\sigma _2/(d\sigma _0)},$$ (67) and defining $$A(x)=\underset{l1}{}\frac{(1)^{l1}}{l^2}x^l=_0^x\frac{dt}{t}\mathrm{ln}(1+t),$$ (68) we find that $$\sigma _0^{}=\frac{\sigma _2}{4d^2A\left(p_c/(1p_c)\right)}(pp_c)+O\left((pp_c)^2\right),$$ (69) and that $$\sigma _{\text{eff}}=\frac{\sigma _2}{4d^2A\left(p_c/(1p_c)\right)}(pp_c)^2+O\left((pp_c)^3\right)(pp_c),$$ (70) where the conductivity exponent is $`t=2`$. Hence, as in the previous “one-impurity” approximation, the replica-coupling ansatz yields critical exponents $`s=0`$, $`t=2`$, and an asymptotic dependence of the threshold $`p_c1/(2d)`$ when $`2d1`$. One can easily check that this new “type 2” effective-medium formula is exact to second order in the weak-contrast limit and in the dilute limit. After a few manipulations, we now obtain with (37) and (59) $$\frac{\mathrm{\Delta }E^2}{E^2}=2q\left(1+\frac{q}{2}\right).$$ (71) ## V Discussion We plot in Fig. 1 (resp. 2) the “type 2” scaled conductivities $`\sigma _{\text{eff}}/\sigma _1`$ versus $`p`$, the volume fraction of material 2, for a dielectric ratio $`\sigma _2/\sigma _1=10`$ (resp. $`\sigma _2/\sigma _1=1000`$). We also show the Hashin-Shtrikman (HS) bounds , the Hori-Yonezawa formula which comes from a cumulant series (CS) approximation, and the one-impurity (OI) Bruggeman formula. The dimension is $`d=2`$. Figs. 3 and 4 display similar plots for $`d=3`$. For moderate contrast (Figs. 1 and 3), we observe that all four self-consistent formulas lie close to each other. This is a consequence of the fact that they are exact to second order in the contrast. Also, for any contrast, the slopes at $`p=0`$ and $`p=1`$ are all identical, which is a consequence of the fact that they are exact to second order in the dilute limit $`p0`$ (the expression near $`p=1`$ is obtained by replacing $`p`$ by $`p1`$ and by interchanging $`\sigma _1`$ and $`\sigma _2`$). We also observe that the HS bounds are satisfied in each case considered. However, the formulas obtained via the cumulant series summation, i.e. both the HY formula and its “type 2” counterpart, do not reduce to the exact result $`\sigma _{\text{eff}}=1/\sigma ^1`$ in dimension 1 (not shown). This exact result is also the common value of the HS bounds for $`d=1`$. Hence, formulas derived from the cumulant series approximation do not obey the HS bounds in dimension $`d=1`$. On the other hand, both the Bruggeman formula and its “type 2” counterpart do reduce to the exact result when $`d=1`$, and can be seen to always obey the HS bounds whatever $`d`$ is. The one-impurity approximation scheme therefore appears to be of better physical relevance for all dimensions, than the cumulant series approximation. We now discuss the critical behavior. First of all, the percolation thresholds found in the“type 2” formulas are $`p_c=1\mathrm{exp}\left(1/(2d)\right)`$ (cumulant series) and $`p_c=1/(2d1)`$ (one-impurity). These thresholds are the percolation thresholds of the Potts model, and that of the Bethe lattice model, respectively. Both thresholds decrease as $`1/(2d)`$ when $`d\mathrm{}`$, which is the exact asymptotics. The Bethe and Potts models are mean-field models, where emphasis is put on fluctuations in the couplings between a given site and its neighbours. On the contrary, in effective-medium theories, interactions between impurities are taked into account through the self-consistent background medium. Such interactions are more important for low dimensions. Effective-medium theories therefore overestimate interactions in high dimensions, whereas mean-field models are expected to underestimate them in low dimensions. Above the upper critical dimension where mean-field models are accurate, interactions between impurities become irrelevant. According to this discussion, our “type 2” formulas appear as hybrids between mean-field and usual effective-medium theories, and are expected to be mostly relevant in dimensions intermediate between $`d=1`$ and the upper critical dimension $`d=6`$. Indeed, the condition $`\sigma _{\text{eff}}^{}(\sigma _0)=0`$ minimizes the influence of the background medium and, according to the interpretation developped in Sec. III B, replica coupling has to do with couplings between neighboring points. The reason for which the introduction of a replica-coupling ansatz yields the exact thresholds of mean-field theories will have to be clarified in the future. In Fig. 5, we plot the quadratic fluctuations $`\mathrm{\Delta }E^2/E^2`$ as a function of $`p`$. The fluctuations in the “type 2” estimates are greatly reduced compared to those of the Bruggeman and HY formula. This is consistent with the fact that the influence of the background is reduced. “Type 1” formulas give exponents $`s=t=1`$, while for “type 2” formulas they are $`s=0`$, $`t=2`$. Mean-field theories yield $`s=0`$, $`t=3`$ which are the exact values for $`d6`$. It is interesting to compare these values to exact bounds deduced from the Nodes-Links-Blobs (NLB) model, in all dimensions. The NLB model is currently accepted as a good one for the backbone structure of real random resistor networks. The bounds read $`t1+(d2)\nu ,`$ (73) $`s1+(2d)\nu ,`$ (74) where $`\nu >0`$ is the correlation length exponent: $`\xi |pp_c|^\nu `$. They hold for $`2d6`$, whereas for $`d>6`$ the right-hand sides in (V) are fixed to their $`d=6`$ values. These bounds follow, e.g., from comparing the lower and upper exact bounds obtained in Ref. for the noise exponent $`\kappa `$ in weakly nonlinear networks, within the NLB scheme. They are satisfied by simulation results . Using the usual effective-medium values $`s=t=1`$ in (V) implies the absurd value $`\nu =0`$, save for $`d=2`$ where a finite value of $`\nu `$ is allowed. Though information about the correlation length $`\xi `$ (and therefore about $`\nu `$) is not included in the Bruggeman nor in the HY formulas, the above bounds show that, as long as they are meant to model percolating systems obeying the NLB picture, these formulas are truly adequate only in dimension $`d=2`$ – and $`d=1`$ where the Bruggeman formula is exact. As to “type 2” formulas, we insert the values $`s=0`$, $`t=2`$ into (V) and deduce that $`\nu =1/(d2)`$, a reasonable expression for $`d3`$ only. If we furthermore insist on having $`\nu 1/2`$ as in real systems, these heuristic arguments restrict the range of validity of the new formulas to $`d=3,4`$. Note, moreover, that only in dimension $`d=2`$ are the exponents equal: $`s=t`$, because of self-duality . A formula with unequal exponents therefore is expected to be essentially relevant to dimensions $`3`$. We also quote theoretical bounds for $`t`$ due to Golden, valid for hierarchical NLB models: $`1t2`$ for $`d=2,3`$ and $`2t3`$ for $`d4`$ . The above analysis is consistent with these bounds, and can be summarized as a set of prescriptions for using the “best” available effective medium theories, as far as a non-conflicting critical behaviour is concerned: for $`d=1`$, Bruggeman’s formula, or its “type 2” counterpart are exact; for $`d=2`$, the Bruggeman or HY formulas are adequate; for $`d=3,4`$, “type 2” formulas are applicable; finally, for $`d5`$ mean-field theories would be the most relevant. Estimates or exact values for the exponents are : $`(\nu ,s,t)=(4/3,1.3,1.3)_{d=2}`$, $`(0.88,0.73,2.00)_{d=3}`$, $`(0.68,0.4,2.4)_{d=4}`$, $`(0.57,0.1,2.7)_{d=5}`$, $`(1/2,0,3)_{d6}`$. These values support our prescriptions. An interesting observation is that actually both “type 1” and “type 2” formulas can be given by a variational formulation as $`\sigma _{\text{eff}}^{\text{type 1}}`$ $`=`$ $`\underset{\genfrac{}{}{0pt}{}{\sigma _00}{0q(\sigma _0)1}}{\mathrm{min}}\sigma _{\text{eff}}(\sigma _0),`$ (75) $`\sigma _{\text{eff}}^{\text{type 2}}`$ $`=`$ $`\underset{\genfrac{}{}{0pt}{}{\sigma _00}{0q(\sigma _0)1}}{\mathrm{max}}\sigma _{\text{eff}}(\sigma _0),`$ (76) provided that an unphysical solution $`\sigma _{\text{eff}}=0`$ is discarded in the minimization (75). Indeed, at least in the framework of the two different models introduced in Sec. IV, the curves for $`q(\sigma _0)`$ and $`\sigma _{\text{eff}}(\sigma _0)`$ are found to have the form shown in Fig. 6. The infimum (75) occurs at $`q=0`$, whereas the solution $`\sigma _0^{}`$ to the equation $`\sigma _{\text{eff}}^{}(\sigma _0)=0`$ corresponds to a maximum of $`\sigma _{\text{eff}}`$. Both types of theories can therefore be interpreted as extremal theories in the framework of self-consistent models built on the replica-coupling ansatz. The physical meaning of this interpretation is still not clear. However, “type 2” formulas should not been disregarded as unphysical because of their showing up as maximal ones: the minimization principle states that the dissipated power is minimized with respect to the electric field; but there is no reason why an extremization with respect to arbitrary variational parameters should not lead to a maximum of the dissipated power. Eqs. (75), (76) explain why for a given approximation, one always has $`\sigma _{\text{eff}}^{\text{type 1}}\sigma _{\text{eff}}^{\text{type 2}}`$ in Figs. 1-4. We now consider some points that were not explicitly treated in the paper. First, we presented the formalism in terms of the electric field $`E`$, from which we obtained a conductivity $`\sigma _{\text{eff}}`$. The electric current $`j`$ (or the induction $`D`$), could be used instead . In such a formulation, the random constitutive parameter is the resistivity $`\rho (x)=1/\sigma (x)`$ and the constraints are $`j=0`$ and $`\overline{j}=j_0`$. One then computes an effective resistivity $`\rho _{\text{eff}}`$. Both formulations are equivalent, but a given approximation scheme in general leads to different results for $`\sigma _{\text{eff}}`$ and $`\stackrel{~}{\sigma }_{\text{eff}}=1/\rho _{\text{eff}}`$. Preliminary investigations of “type 2” formulas have been led in this case. These will be presented elsewhere. Finally, we discuss the natural question about the possibility of replica-symmetry breaking . Replica symmetry breaking introduces more free parameters in the ansatz, and the final extremization has to be carried out with respect to several variables. There is no frustration in this problem, and we therefore expect the replica symmetric solution to be the only one. In order to test this, we tried a one-step symmetry-breaking solution and did indeed not find any new solution. ## VI Conclusion We presented a functional approach to the calculation of effective-medium properties of random media. We showed how to recover the Bruggeman and Hori-Yonezawa formulas by using specific approximation schemes to the basic functional integral. We also discussed the introduction of a replica-coupling parameter in a gaussian Ansatz, from which new effective-medium formulas were obtained. These formulas appear to be more adequate in $`d=3`$ compared to the standard ones by Bruggeman and HY. Because it yields a sensible result in all dimensions, and fulfills all the constraints required to deserve the label of a “good” effective-medium theory, the “type 2” counterpart of the Bruggeman formula offers an interesting alternative to the latter. Indeed, it has a percolation threshold equal to $`p_c=1/5`$ in three dimensions. This is closer to values observed in real materials, compared to the $`p_c=1/3`$ of the Bruggeman formula which often constitutes an overestimation. ## ACKNOWLEDGMENTS We gratefully acknowledge H. Orland for stimulating discussions. One of us (MB) wants to thank H.E. Stanley for his hospitality at the CPS and the DGA for financial support. ## A Quadratic fluctuations of the field For completeness, we give here the demonstration of Eq. (37) . Volume averages are identified to statistical ones. Because $`W^{}=(1/2)\sigma E^2=(1/2)_\nu p_\nu \sigma _\nu E^2_\nu `$, where $`_\nu `$ denotes an average on the phase $`\nu `$ on which the conductivity $`\sigma _\nu `$ is constant, the effective conductivity reads $$\sigma _{\text{eff}}=\underset{\nu }{}p_\nu \sigma _\nu \frac{E^2_\nu }{E_0^2}.$$ (A1) On the other hand, $`\sigma _{\text{eff}}`$ has to be an homogeneous function of degree one of the $`\sigma _\nu `$, whence $$\sigma _{\text{eff}}=\underset{\nu }{}\sigma _\nu \frac{\sigma _{\text{eff}}}{\sigma _\nu }.$$ (A2) Comparing both equations yields the values of the $`E^2_\nu `$ and consequently that of $`E^2=_\nu p_\nu E^2_\nu `$. Equ. (37) follows. We note that if $`\sigma _{\text{eff}}=\sigma `$ (an exact upper bound for the effective conductivity), then $`\mathrm{\Delta }E^2=0`$. Therefore, $`\sigma _{\text{eff}}=\sigma `$ defines a trivial model of a medium which is a composite, but from which field fluctations are nonetheless absent. ## B Calculations in the “one-impurity” approximation The approximation which leads to (42) is built as follows. We first expand the exponential $$e^{\beta (_e_0)}_0=\underset{k0}{}\frac{(\beta )^k}{k!}(_e_0)^k_0.$$ (B1) Since $`_0`$ is non-random, using the hamiltonian density $`w_0`$ defined in (40) and $`\mathrm{\Delta }w_x[E(x)]`$ defined by (41) we can rewrite the difference $`_e_0`$ as $$_e_0=𝑑x\mathrm{\Delta }H(x),$$ (B2) where $$\mathrm{\Delta }H(x)=\frac{1}{\beta v}\mathrm{ln}e^{\beta v\mathrm{\Delta }w_x[E(x)]}.$$ (B3) The one-impurity approximation consists in writing ($`k1`$) $`(_e_0)^k`$ $`=`$ $`{\displaystyle 𝑑x_1\mathrm{}𝑑x_k\mathrm{\Delta }H(x_1)\mathrm{}\mathrm{\Delta }H(x_k)}`$ (B5) $``$ $`v^{k1}{\displaystyle 𝑑y\mathrm{\Delta }H(y)^k}.`$ (B6) The last expression only retains contributions from identical points in Eq. (B5). Summing back the series in (B1), and using (B3) yields $`e^{\beta (_e_0)}_01+{\displaystyle \frac{dy}{v}(y)1},`$ (B7) $`(y)={\displaystyle \frac{\stackrel{~}{𝒟}(E^\alpha )e^{\beta {\scriptscriptstyle 𝑑x\left\{w_0[E(x)]+v\mathrm{\Delta }w_𝐱[E(x)]\delta (xy)\right\}}}}{\stackrel{~}{𝒟}(E^\alpha )e^{\beta _0}}},`$ (B8) where a one-impurity-type integral is involved. Since the fundamental size of the theory ($`v^{1/d}`$) is much smaller than the volume $`V=1`$ of the system, and since the latter is statistically translation-invariant, the outer integral over $`y`$ is redundant with the disorder average, and can be dropped. We therefore arrive at (42). When $`\beta \mathrm{}`$, the functional $`(y)`$ can be computed exactly. Let us briefly indicate how to do it. We first introduce the notation $`\stackrel{}{h}`$ for vectors of dimension $`nd`$, and components $`h_i^\alpha `$, with $`\alpha =1,\mathrm{},n`$, $`i=1,\mathrm{},d`$. Hence, $`\mathrm{\Delta }w_y[E(y)]\mathrm{\Delta }w_y\left(\stackrel{}{E}(y)\right)`$. The next step is to use the formal identity $$e^{\beta v\mathrm{\Delta }w_y(\stackrel{}{E})}=\frac{d\stackrel{}{h}d\stackrel{}{h}^{}}{(2\pi )^{nd}}e^{i\stackrel{}{h}\stackrel{}{h}^{}}e^{\beta v\mathrm{\Delta }w_y\left(i\frac{}{\stackrel{}{h}^{}}\right)}e^{i\stackrel{}{h}^{}\stackrel{}{E}}$$ (B9) to write the numerator of (B8), which we denote hereafter by $`𝒥(y)`$, as $$𝒥(y)=\frac{d\stackrel{}{h}d\stackrel{}{h}^{}}{(2\pi )^{nd}}e^{i\stackrel{}{h}\stackrel{}{h}^{}}e^{\beta v\mathrm{\Delta }w_y\left(i\frac{}{\stackrel{}{h}^{}}\right)}\stackrel{~}{𝒟}Ee^{(\beta /2){\scriptscriptstyle 𝑑x\stackrel{}{E}(x)\stackrel{~}{M}\stackrel{}{E}(x)}+i\stackrel{}{h}^{}\stackrel{}{E}(y)},$$ (B10) where $`\stackrel{~}{M}`$ is the matrix defined from the replica-coupling matrix $`M`$ in $`w_0`$ by $`\stackrel{~}{M}_{ij}^{\alpha \gamma }M^{\alpha \gamma }\delta _{ij}`$. After an integration over the fields $`E`$ and $`\varphi `$ implied in the measure $`\stackrel{~}{𝒟}E`$ (which can be easily done using the Fourier components of $`\varphi `$, and with $`y=0`$ since (42) is independent of $`y`$), $`𝒥`$ reads, up to inessential factors : $$𝒥(y)=e^{(\beta /2)_{\alpha \gamma }M^{\alpha \gamma }E_0^2}\frac{d\stackrel{}{h}d\stackrel{}{h}^{}}{(2\pi )^{nd}}e^{i\stackrel{}{h}\stackrel{}{h}^{}}e^{v\beta \mathrm{\Delta }w_y\left(i\frac{}{\stackrel{}{h}^{}}\right)}e^{i\stackrel{}{h}^{}\stackrel{}{E}_0\stackrel{}{h}^{}\stackrel{~}{M}^1\stackrel{}{h}^{}/(2\beta vd)}.$$ (B11) Formally expanding $`\mathrm{exp}(v\beta \mathrm{\Delta }w_y)`$ in powers of $`i/\stackrel{}{h}^{}`$, and carrying out successive integrations by parts over $`\stackrel{}{h}^{}`$ yields $$𝒥(y)=\left[\text{Det}(M)\left(\frac{v\beta d}{2\pi }\right)^n\right]^{d/2}e^{(\beta /2)_{\alpha \gamma }M^{\alpha \gamma }E_0^2}𝑑\stackrel{}{h}e^{(v\beta d/2)(\stackrel{}{E}_0\stackrel{}{h})\stackrel{~}{M}(\stackrel{}{E}_0\stackrel{}{h})v\beta \mathrm{\Delta }w_y(\stackrel{}{h})},$$ (B12) where the determinant is evaluated in replica space. Finally, $`(y)=𝒥(y)/𝒥(y;\mathrm{\Delta }w_y=0)`$: $$(y)=\left[\text{Det}(M)\left(\frac{v\beta d}{2\pi }\right)^n\right]^{d/2}\underset{\alpha }{}dh^\alpha e^{v\beta \left\{\frac{d}{2}_{\alpha \gamma }M^{\alpha \gamma }(E_0h^\alpha )_i(E_0h^\gamma )_i+\mathrm{\Delta }w_y[h]\right\}}.$$ (B13) For any $`\mathrm{\Delta }w_y`$, this integral over the replicated vector field $`h`$ can be computed exactly using a saddle-point method in the limit $`\beta \mathrm{}`$, as announced. This allows for a possible extension of the theory to nonlinear media in the “one-impurity” approximation. Here, for the linear problem at hand, Eq. (B13) is a simple gaussian integral. Setting $`\mathrm{\Delta }\sigma =\sigma \sigma _0`$ and $$\mu =\left(1+\frac{\mathrm{\Delta }\sigma }{d\sigma _0}\right)^1,$$ (B14) we obtain $$\mathrm{ln}(y)=\left[\frac{\beta v}{2}\mathrm{\Delta }\sigma \mu E_0^2+\frac{d}{2}(\mathrm{ln}\mu \beta vq\sigma \mu E_0^2/d)\right]n+O(n^2),$$ (B15) from which follows (44). ## C Calculation in the “cumulant series” approximation We have to compute $`_e_0`$ and $`_0_0`$ in Equ. (56). The calculation of $`_0_0`$ is easy with the methods already employed, and yields $$\underset{n0}{lim}_0_0/n=\frac{1}{2}\sigma _0E_0^2.$$ (C1) As to $`_e_0`$, we first expand $`_e`$ (Eq. (13)) in the cumulants $`C_k`$ of the disorder averages of $`\sigma (x)`$, according to their definition by the generating funtion ($`X`$ is a generic expansion variable) $$\mathrm{ln}e^{X\sigma }=\underset{k1}{}\frac{X^k}{k!}C_k(\sigma ).$$ (C2) We therefore have: $$_e=\frac{1}{\beta }\underset{𝐱}{}\underset{k1}{}\frac{1}{k!}\left[\frac{\beta v}{2}\underset{\alpha }{}E^\alpha (𝐱)^2\right]^kC_k(\sigma ).$$ (C3) We deduce that $$\frac{1}{n}_e_0=\frac{1}{V\beta }\underset{𝐱}{}\underset{k1}{}\frac{(\beta v)^k}{k!}C_k(\sigma )𝒞_k(E^2/2),$$ (C4) where $$𝒞_k(E^2/2)=\frac{1}{n}\left[\underset{\alpha }{}E^\alpha (x)^2/2\right]^k_0$$ (C5) (because of statistical homogeneity, these coefficients do not depend on the position variable $`x`$). It is convenient to introduce the following generating function $`𝒵(X)`$ in order to compute the $`𝒞_k`$: $`𝒵(X)`$ $`=`$ $`{\displaystyle \underset{k1}{}}{\displaystyle \frac{(X)^k}{k!}}𝒞_k(E^2/2)`$ (C6) $`=`$ $`{\displaystyle \frac{1}{n}}\left\{\mathrm{exp}\left[{\displaystyle \frac{1}{2}}X{\displaystyle \underset{\alpha }{}}E_{}^{\alpha }{}_{}{}^{2}(x)\right]_01\right\}`$ (C7) $`=`$ $`{\displaystyle \frac{1}{n}}\mathrm{ln}\mathrm{exp}[{\displaystyle \frac{1}{2}}X{\displaystyle \underset{\alpha }{}}E_{}^{\alpha }{}_{}{}^{2}(x)]_0+O(n).`$ (C8) Setting $`A^{\alpha \gamma }=\delta _{\alpha \gamma }+(X/v\beta d)[M^1]^{\alpha \gamma }`$, we obtain $$𝒵(X)=\frac{1}{2n}\left\{d\text{Tr}\text{Ln}A+XE_0^2\underset{\alpha \gamma }{}[A^1]^{\alpha \gamma }\right\}+O(n)$$ (C9) (the trace and the logarithm act in the replica space). Expanding (C9) in powers of $`X`$ then allows for the identification $$𝒞_k(E^2/2)=\frac{k!}{2}\left(\frac{1}{v\beta d}\right)^k\frac{d}{n}\left[\frac{1}{k}\text{Tr}(M^k)+v\beta dE_0^2\underset{\alpha \gamma }{}[M^{1k}]^{\alpha \gamma }\right]+O(n).$$ (C10) Use of this expression in (C4) cancels the convergence factor $`k!`$: we reintroduce it by inserting the identity $$\frac{1}{m!}_0^{\mathrm{}}𝑑ue^uu^m=1,$$ (C11) applied to $`m=k1`$ and $`m=k`$ in the resulting cumulant series. This permits its Borel summation, which brings in the functions $`h_m(x)`$ defined by (C13). This results in $$\frac{1}{n}_e_0=\frac{1}{n}\left[\frac{d}{2}E_0^2\underset{\alpha \gamma }{}\left[Mh_0(M^1/d)\right]^{\alpha \gamma }+\frac{1}{2v\beta }\text{Tr}h_1(M^1/d)\right]+O(n),$$ (C12) where we defined the family of functions $$h_m(z)=_0^{\mathrm{}}𝑑uu^me^u\mathrm{ln}e^{u\sigma z}(m>2).$$ (C13) Note for further use that $$h_m^{}(z)=\frac{1}{z}\left[h_{m+1}(z)(k+1)h_m(z)\right].$$ (C14) For $`M`$ given by (30), the differents terms in (C12) are $`\underset{n0}{lim}{\displaystyle \frac{1}{n}}{\displaystyle \underset{\alpha \gamma }{}}\left[Mh_0(M^1/d)\right]^{\alpha \gamma }=\sigma _0h_0\left(1/(d\sigma _0)\right),`$ (C15) $`\underset{n0}{lim}{\displaystyle \frac{1}{n}}\text{Tr}h_1(M^1/d)=dh_1\left(1/(d\sigma _0)\right)+v\beta d\sigma _0qh_0\left(1/(d\sigma _0)\right)E_0^2.`$ (C16) which leads to Eq. (57).